and u and the unknown coefficients a and ~ using the calculus of variations. However, the set of governing equations so derived, which consists of four simultaneous firstor second-order partial differential equations, is rather difficult to solve. To make the solution more tractable, the following SiOlplifications are intro-
duced. Approximate Solution Method Minimization of the total potential energy with respect to , which· may be interpreted as the moment equilibrium equation and expressed in the following form: £1 o = M dz 1227
(11 )
where El = effective bending stiffness of the tubular structure and M is the overturning moment of the lateral load. This equation is not easy to solve because £1 varies with height and is dependent on other unknowns. Nevertheless, if the effect of the variation of EI with height on the bending rotation is assumed negligible, then
l
z
=S
(13)
in which S = shear of the lateral load. This is actually the horizontal shear equilibrium equation. From this equation, u can be determined by direct integration as follows:
u
= (z ( _ S -
Jo
4G w twQ
<1»
dz
(14)
~
and
(X
l3
(2)
(3)
Puint luatl.at
= , 1.17nl". +
l.OU
QI
=
+
1.00
(X"
Q,
lOp
-
Uniform distributed load
= 111;
+ 2.67m", + 0.57
~~
= HI} + 11.20n" + Hl.OR
2.57m no + 1.12 = Ill;'. + 2.94/11". + 0.04
~.
=
+ 2.6711'". + U.S7
O.29In", m~.
O.03m". + 1.12 ex" = - - - - - - - - - Triangular distributed load
(X
-
m;. + 2.94m", + 0.64
,
+ 1.09 = m~. 2.22m". + 2.86,nt\' + 9.62
OJ
J.SOn" + 12.60 + tl.:!()I1'r + to.OX
~I
111 ~,.
= m;'.
O.10m", + 1.09 + 2.86111", + 0.62
O.RSln,. + 11.00
nl i,
7.721n, + 14.15 15111, + II ..l -' + I'- ... O.ORnl,. + 14.15 II ._ 3' 1.
~~ = rtl;. + I'_.."5 In, + J3. =
J3
6.67/111' + 13.71 . nlj + 12.01111{ + 10.97 7 __0_.2_9_11...:..'1_+_13_._ In} + 11.01",{ + 10.97 _1_
~
=
_ G w H2 -~ E w a2
(17)
G H2 m / / - E b'1
( 18)
m H'
Substituting the preceding values of
Q
(1 )
0
(~~ +
Formulas for
Load case
(12)
M dz
Likewise, minimization of the total potential energy with respect to u yields the following governing equation for u: 4G w t,.tl
TABLE 1.
/
It should be noted from the formulas given in the table that the shear lag coefficient of a frame panel is dependent only on the elastic properties of that particular panel, not on those of any other panels. EFFECTS OF VARIOUS PARAMETERS ON SHEAR LAG
where a .. (X2' f3 .. and f32 = unknown coefficients to be solved. Note that al and f31 are actually the values of (X and f3 at the base, while (X2 and f32 are the corresponding values at the top. Substituting the foregoing equations into the expression for the total potential energy, minimizing with respect to a. (X2' f3., and f32' and solving the set of algebraic equations thus obtained, the shear lag coefficients are determined for each loading case. The results are given in Table 1 in which the relative shear stiffness parameters m w and In, are defined by
The shear lag coefficients are plotted against the relative shear stiffness parameters of the frame panels for each loading case in Fig. 5. Fronl this figure, the effects of the various parameters on the degrees of shear lag are readily revealed. Firstly, it can be seen that in all cases, a. > a2 and f3. > f32 and therefore, the shear lag effects are generally greater at the base of the structu re than at higher levels. Secondly, it is apparent that the degrees of shear lag are dependent on the distribution of lateral loads. At the base of the structure, the degrees of shear lag increase with the following order of load cases: point load at top, triangularly distributed load and uniformly distributed load. However, at the top of the structure, the degrees of shear lag decrease wi th the same order of load cases. Thirdly, since the shear lag coefficients decrease with the relative shear stiffness parameters of the panels, the shear lag effects may be reduced by increasing the frame member sizes. Moreover, as the shear lag effects are largest when the relative shear stiffness parameters, which are proportional to the square of the height/width ratios of the frame panels, are small, the shear lag effects are generally more significant in low-rise buildings than in high-rise buildings. Lastly, it is identified that the single most important structural parameter
1228
1229
(1 - ~Y + [2 ~ - (~y] ~ = ~I (1 - ~r ~ ~2 [2~ - (~y]
a = a.
a2
t
(15)
(16)
flange panels can be evaluated by first differentiating the axial displacements with respect toz as per (3)-(4), and then multiplying the axial strains so determined bytheir respectively Young's moduli. The expressions thus obtained for the axial stresses generally consist of two terms, one proportional to d4>/dz and the other proportional to 4>. At the base of the structure, where the axial stresses are most critical, cP is equal to zero and the expressions for the axial stresses would be reduced to
load case
-.
---. ___ •
-- --.--j}
«1
2
5
4
3 m
w
1.0
load case
N
QJ. L-
0
-.-
0.8
1 2
ct:J.....
3
0.6
- - }~I - --.. .
0.4
0.2
0
0
} f3 2
1
4
3
5
~
that determines the degree of shear lag in a frame panel is the relative shear stiffness parameter of the panel as· defined by (17) or (18). The effects of changing the sizes of the frame members can be evaluated simply by calculating the new values of relative shear stiffness parameters and using the design chart given in Fig. 5 to determine the corresponding values of shcur I,lg coefficients. This nlethod is, therefore, particularly suitable for preliminary evaluations of the main structural element sizes during the early stages of design. STRESSES AND DEFLECTIONS
Axial Stresses
Having determined the shear lag coefficients ex and f3 and hence the distributions of the axial displacements, the axial stresses in the web and 1230
(20)
I:n 2t CTzX dx w
+
fb
2tf CT;a dy
+ 4A k CT k a
(21 )
= M
d/dzis determined as dcf>/dz = MIEI, where EI is given by
= ~ Ewtwa)
(1 - ~ ex) + 4E
f lf
a 2b
(1 - ~ f3) + 4E",A a k
2
(22)
2
r
Values of a and
(19)
At levels above the base, however,
EI
m FIG. 5.
= Ew ~~ a f
)«
2
[(1 - ex); + ex (;rJ CT; = E ~~ a [(1 - ~) + ~ (~r]
az
Putting the value of d4>/dz so obtained back into (19)- (20). the axial stresses can be expressed directly in terms of the overturning nlonlent. Lateral Deflections The lateral deflection u of the structure can be evaluated by first substituting the value of EI into (12) to solve for thus obtained into (14) to solve for u. Since the value of £1 varies with height, the resulting expressions for
(1
u= -P - H Z 2 EI 2
-
1) + -4G- -
-
6
1231
zJ
P
w l wQ
z
(23)
Load case 2-uniform load
(1
c:
1
1) + - -
u = -U - H2 z2 - - HzJ + £1 4
6
24
Z4
U
4G w tw a
1)
( Hz - 2
Z2
.2
U
.!
(24)
~
c
~ C .2
Load case 3- triangular load u = -T
(1 H2
EI 6
Z2 -
(1
1 H z3 + - 1 -ZS) + -T- - Hz - -1 -Z3) 12 120 H 40 w t..,a 2 6H
-
-g
o o....
(1)"8
(25)
8..£:
o·~
0.&
~
Example 1 A high-rise 40-story reinforced concrete framed tube structure, as shown in Fig. 6(a), is analyzed. All the beam and column members are of sizes 0.8 m x 0.8 m. The height of each story is 3.0 m and the center-to-center spacings of the columns are 2.5 m. The Young's and shear moduli of the material are 20 GPa and 8.0 OPa, respectively. A uniformly distributed lateral load of 120 kN/m is applied to the structure. The equivalent elastic properties of the analogous orthotropic membrane tube, as evaluated by the method given in Appendix I, are as follows:
!l
=~
III 61>' L. Q
+J
....
(
' ....
u
oII)
......... "
~
=
0, = 1.441 GPa
= 0.256 m
t w = I,
Ak
=0
(ij
·x
< ~
-0 OJ ~
c:
< OJ :;
(I)
~----,
U;~ ~
&1-4
eu eu &.. I:: \to4 eu
..0 )
(27)
+J
.£:
,- 7
."
I
N
(29)
.c:I
"0
..-4
GJ
~ ns
1-4
~
en Q)
GJ
.!.H
2
QJ
:r.
bO
~
U
.D
'/
..-4
From these elastic properties, the relative shear stiffness parameters pf the web and flange panels are worked out by (17)-(18) as
en
."
~
."
....
~
..0
-tJ
& ~
'"
:0
GJ
"0
o
0
o...c: 0 .....
... en
~
Q)
m,
= 3.388
(30a,b)
.0
o.e
1::lJ.I-J.L...J
...OJ
t>:::s
bO C co -e
GJ (/1"0
~ :s
co
&.. GJ
m.., = 4.611;
... en
N
(26)
(28)
en en Q)
n;
GJ
OK'
u
~o 0
~
Q)
Ew = £, = 20.0 GPa
Q)
-t ~
"0
~
......
~ns
~, Jl=
en C
QJ
' ... ""
.£:
.0 .t:
c ~
I.. GJ
EXAMPLES AND COMPARISONS WITH COMPUTER ANALYSIS
:;
e
&
....:::J
~
"C C1)
The shear lag coefficients are then determined by the formulas given in
E
co
Table 1 as
u: a, = 0.366;
Ct2
= 0.624;
~2
~l
= 0.035 = 0.223
OJ
(31a,b)
C/)
(32a,b)
Having determined the shear lag coefficients, the bending stresses at the base and at mid-height of the structure are evaluated by (19)-(20) and plotted in Fig. 6(b). The deflections of the structure are then calculated by (24) and plotted in Fig. 6(c). For comparison, the corresponding results obtained by using a standard space-frame analysis program are also plotted alongside the foregoing results. It is revealed that the proposed method underestimates the maximum axial stress at the base by 14% and overestimates the maximum lateral deflection by 13%.
_ ~
a:,
C'l .t:
i: '0 C/)
'iii >
'i r:
/
r
/
7
w Oll
< .,..I C1)
0..
E CO
)(
Example 2 In this example, a low-rise IS-story framed tube structure constructed of structural steel, Fig. 7(a), is analyzed. All the beam and column members arc 610 x 305 x 238 kg/n, Universal Benrns (I = 207,571 cnl"; A = 303.R C1l1 1 ; A. = 117.7 cm 2). The height of each story is 3.2 m and the center to
UJ
cD
l!o
f
1232
C1)
· :::Jc:
C)
1233
c
center spacings of the columns are 2.8 ffi. The Young's and shear moduli of the material are 200 OPa anti 80 OPa, respectively. A triangularly distributed lateral load of intensity 150 kN/m at the top and zero intensity at the bottom is applied to the structure. The equivalent elastic properties of the analogous orthotropic membrane tube are determined according to Appendix I as follows:
o
n Q)
~
o
..-.. ~
0
M
....
C .2 '5 .a .t:
! ~
~
U GJ
.... ....
(33)
U)
Ak = 0
(36)
ca ~
From these equivalent properties, the relative shear stiffness parameters are calculated as
1;)
d 0
-e
""
= 200 OPa G w = G, = 5.345 GPa I ... = If = 0.0109 m EM' = E1
C en CI)
...
..... ..... QJ
(34) (35)
Q)
"C
e
mK' = 0.218;
"C
m/
= 0.314
(37a,b)
Q)
-~
...
~
and the corresponding shear lag coefficients are obtained as:
coc
~
<
...
)
.tJ
..... .t:
QJ
GJ
..0
"0
-e
GJ
tl()
C
ns
S o6.J
ns
-~-:
~
"C GJ
...
CX2
= 0.861
(38a,b)
(31
1.065;
132
= 0.930
(39a ,b)
RS
-..0
~-g
Figs. 7(b and c) illustrate the final results for the axial stresses and lateral deflections. The corresponding frame analysis results are also plotted in the above figures "to demonstrate the accuracy of the proposed method. I t is seen from the comparison that the proposed method underestimates the maximum axial stress at the base by 5% and overestimates the maximum lateral deflection by 8%. These errors are quite acceptable for preliminary design purposes. However, since the values of a and ~ at the base of the structure are both greater than 1.0, the proposed method yields axial stresses of the wrong sign near the centroidal axis of the frame panels at the base of the structure. Fortunately, as the axial stresses there are actually very small, such errors are not of much practical importance.
en
..-.. .!,
...:JQ)
o...c
1j
c.e
..... en
...:J
Q+J L. GJ
~
Jl:J .... "C (1)
E ~ u. ~
CONCLUSIONS
a:
-nt
~ o
...I
....o
.u;U) ~
"i
c
<
I
C'I
r
1.129;
:J C'3
I
~
1j en
.t::
RS
:J
:&
t:IO
-e GJ
/
al
(1)
.tJ 4J
I
w 8~
I
Go)
0. E n) ><
w r-=
.
Q)
~
CJ :::s
i!o
1234
A simple hand calculation method for approximate analysis of framed tube structures with the shear lag effects taken into account was proposed . In the proposed method, independent distributions of axial displacCIl1Cnts are used for the web and flange panels and thus the shear lag in each panel is individually allowed for. Closed-form solutions are obtained, from which the effects of various parameters on the overall structural behavior can be readily evaluated. The single most important parameter that determines the degree of shear lag in a frame,panel is the relative shear stiffness parameter of the panel as defined by (17) or (18). Since the shear lag effects of each frame panel can be separately evaluated and the degree of shear lag in a frame panel is dependent only on the properties of that particular panel, the shear lag effects may be calculated by means of only a small set of design charts (Fig. 5). The proposed method is easy to use and yet reasonably accurate and is thus most suitable for preliminary design calculations. Numerical examples demonstrated that the method is applicable to framed tube structures over a wide range of building heights. 1235
Ii
APPENDIX I. ELASTIC PROPERTIES OF eQUIVALENT ORTHOTROPIC MEMBRANES
A typical frame segment bounded by the centers of the adjacent frame members, Fig. 8, constitute a basic unit of the frame and may be modeled 'as a solid membrane spanning the same area (shown by dotted lines in Fig. 8) provided the elastic properties of the membrane are so chosen to represent the axial and shear behavior of the actual framework. The method for evaluating the equivalent properties of the membrane is presented in the following. This method is applicable to both the web and flange panels and thus in the following, they are not distinguished from each other. When it is necessary, the corresponding properties may be denoted with a subscript w to signify their belonging to a web panel, or with a subscript f to signify their belonging to a flange panel.
h./2
h/2
Axial Stiffness Under the action of vertical axial forces, the load-deformation relationships for both the frame unit and the equivalent membrane will be equal if
Est = EmA e
(40)
where E = equivalent elastic modulus of the membrane; t= thickness of the membrane; Em = elastic modulus of the construction material; and A e = sectional area of the column. It is normal practice to fix the value of t such that the area of the membrane is equal to the sectional area of the column (Le. st = A(") and so that the axial stress in the column and that in the membrane are equal. In such a case t
= Ae
(41)
s
centre of column
r
---.,
I
column
h
l
centre of beam beam
I
I I
I
~
boundary of
I
menlbrane
equivalent
L_
.J
~ c
FIG. 8.
Membrane Analogy for Basic Frame Unit 1236
and
E
= Em
(42)
Shear Stiffness Consider now the case of the frame unit subject to a lateral force Q, Fig. 9. The lateral deflection may be computed as the sum of that due to bending 6 h and due to shear 6.s • The bending deflection Ah is given by
Q
I
dbI
Basic Frame Unit under Lateral Shear Force
6. h
I
I I
FIG. 9.
= (It
- db )3 12Em I("
+ (~) 2 (s - dr)'~} s
12E",lb
(43)
where Ib and 1(" = moments of inertia of the beam and column respectively. On the other hand, the shear deflection As is given by ., D. s db) (s - de) (44) Q = OmAs(" + ; G",A sb
(h -
(h) -
in which A sb and A s(" = effective shear areas of the beam and column respectively; and Om = shear modulus of the material. Equating the total lateral deflection of the frame unit to the shear deflection of the membrane, the following equation is obtained: h
Q OS!
= Ab + 6.s
(45)
where G = equivalent shear modulus of the membrane. From this equation, the value of G is derived as 1237
h G
=
st db
mw,m,
(46)
As
-+Q Q
t
Wt
I,
U W
w' a
in which /lJQand AJQ are as given by (43) and (44), respectively. APPENDIX II.
S
f3
REFERENCES
'Yxz
Chan, P. C. K., Tso, W. K., and Heidebrecht, A. C. (1974). "Effect of normal frames on shear walls." Building Sci., Vol. 9, 197-209. Cheung, Y. K. (1983). "Chapter 38: Tall buildings 2." Handbook of structural concrete, F. K. Kong et aI., eds., Pitman Books Ltd., London, England. Coull. A .• and Ahmed, A. A. (1978). "Deflections of frame-tube structures." J. Strllct. Div., ASCE, 104(5),857-862. Coull, A., and Bose, B. (1975). "Simplified analysis of frame-tube structures. J. Struct. Div., ASCE, 101(1 1),2223-2240. Coull. A" and Bose, B. (1976). "Torsion of frame-tube structures." J. Struct. Div., ASCE. 102(12),2366-2370. Coull, A., and Bose, B. (1977). "Discussion of 'Simplified analysis of frame-.tube structures. tt, J. Siruct. Div., ASCE, 103(1),297-299. Coull, A., and Subedi, N. K. (1971). "Framed-tube structures for high-rise buildings." J. Struct. Div., ASCE, 97(8), 2097-2105. Ha, K. H., Fazio, p.. and Moselhi, O. (1978). "Orthotropic membrane for tall building analysis." J. Struct. Div., ASCE, 104(9), 1495-1505. Khan, A. H., and Stafford Smith, B. (1976). "A simple method of analysis :or deflection and stresses in wall-frame structures." Building and En vir. , Vol. 11, tI
1yz
£k Ez E'z
ak az a'1
relative shear stiffness parameter's defined by (17) and (18);
= shear of lateral load;
equivalent thickness of web and flange panels, respectively; lateral deflection of structure; = axial displacement in web panel; axial displacement in flange panel; = shear lag coefficient of web panel; shear lag coefficient of flange panel; = shear strain in· web panel; shear strain in flange panel; = axial strain in corner column; axial strain in web panel; axial strain in flange panel; axial stress in corner column; axial stress in web panel; axial stress in flange panel; and rotation of plane section joining corners of tube.
=
69-78.
Khan, F. R. (1967). "Current trends in concrete high-rise buildings." Proc.. Symp. on Tall Buildings, Coull and Stafford Smith, eds., Pergamon Press, Oxford, England, 571-590. Khan, F. R. (1985). "Tubular structures for tall buildings. Handbook of concrete engineering, M. Fintel, ed., Van Nostrand Reinhold, New York, N.Y., 399-410. Khan, F. R., and Amin, N. R. (1973). "Analysis and design of frame tube struct,u:--es for tall concrete buildings." Struct. Eng., 51(3), 85-92. Spires. D., and Arora, J. S. (1990). "Optimal design of tall RC-framed tube building;." J. Strllct. Engrg., ASCE, 116(4), 877-897. Wong, C. H., EI Nimeiri, M. M., and Tang, J. W. (1981). "Preliminary unul}ksis and member sizing of tall tubular steel buildings. ,. Engrg. J., (Second Quarter), 33-47. It
APPENDIX III.
NOTATION
The following symbols are used in tfzispaper.·
Ak
sectional area of corner column; half width of web panel; half width of flange panel; Em = Young's modulus of material; equivalent Young's moduli of web and flange panels, respec£'1" £, tively; bending stiffness of structure; £1 shear modulus of material; Gm equivalent shear moduli of web and flange panels, respective:y; G"" G, overturning moment of lateral load; M
a b
1238
1?~A
8 Outrigger braced structures
40
......
l~
'j
SP
L
63-21
,j
Behavior of Multi-Outrigger Braced Tall Bui.lding Structures By B. Stafford Smith and I. O. Nwaka
.I:i.~ .j'
{
l I
Synopsis: A study iR made of the forces nnd displacements in multi-outrigger tnll huilding structures! SifTlplified r,eneral equations are developed for the restraining moment of the outriggers, the reduction in drift nnd the optimum location of the outriggers for maximum drift reduction. The efficiencies of various optimum and evenly spaced outriAger systems are presented. The assumptions used make the method of analysis suitable only for preliminary design guidance; however, some v,lluable general conclusions relating to the number and location of outriggers are
'j::; :11,.\
Ii:Ii I
:\:
.11; 'I~
III
1;; ,ill
T
i:i
drawn.
1'\
.,\\
m 11 1
Keywords: bending moments;brac;ng; columns (supports); highrise buildings; lateral pressure; loads (forces); mathematical models; structural analysis. I
~
itl
\:~
.:;~
jll
'It: ",i,
.).J'II
~
JI!;,1.,' ,II
. ~ll ,I'
"'d~1
j~
JI ,\I,'j
1'1 ! I
,I,
":If ~
I
,I
I
1
I
A
r.:1r.:
JIU
t,J La I • V
I U
.., • • ••
•
.....
MU1{I·UUlrlt;O~r ~ll U\...lUI \";~
Bryan Stafford Smith is a Professor of Civil Engineering at McGill University, Montreal. Previously he was a Professor of Civil Engineering at the University of Surrey in England. He has researched over a period of many years on various problems relating to tall building analysis.
acting about a common neutral axis and contributing to the total res~sting moment an amount
fc
1 1
{
,
,i INTRODUCTION Outrigger bracing is an efficient means of reducing the drift and forces in a tall building structure. In its simplest form the system consists of a reinforced concrete or braced steel core to which horizontal cantilevers - outriggers - are rigidly attached at one or mora levels, (Fig. 1). The ends of the outriggers connect to columns which, when the building is subjected to sideways loading, resist the rotation of the outriggers and core. In modifying the free deformation of the core, the total drift of the building and the moments in the core are reduced, (Fig. 2). If the principal restraining columns are on the face of the building it is sometimes expedient to mobili~e the axial stiffness of additional perimeter columns by running a deep, very stiff spandrel girder around the structure at the outrigger levels. If this is placed at the top, the system is sometimes called "hat" or "top-hat" br.acing, and if located at int ermed iate levels, "belt-bracing". Notable recent examples of belt-braced buildings include the 42 - storey First ~~isconsin Centre in 1.1ilwaukee, (Fig. 3), by Skidmore, Owings and Merrill of Chicago, and the 43 - storey Yasuda Insurance Company Building in Tokyn, (Fig. 4) designed by the Structural Department of the Tasei Corporation.
A
Taranath (1) who showed that the level of the outrigger for minimum drift is close to the mid-height of the building. ~tc~abb and Muvdi have also studied the problem, confirminA Taranath's results for a single outrigger (2) and extending their considerations to two outriggers (3). Further reductions in the total drift and the core bending moments can be achieved by increasing the size Rnd therefore nxi:ll Atiffnf'fHl of the' Q(ll\1mn~. ilnd hv lldd ln~ ou t r 19ge rH n t more leve 1M. The lmprovcmcn t d 1m1n 1 ::;hl"fi , however, for each additional level. Taken to the limit. c.)utrigp,ers plnccd at nn infinite number of levels, (FiA. 5) ,,,...\\lld r '" \l ~ l' the r 01 \I mn H t (\ he hnve f \l 11yenmr 0 ~ 1r l' 1~' \\' l t h t \ h" \- \"1 r l' •
(
c
EI + AE
c
d 2/2
.) . N
total
( 1)
in which EI is the flexural rigidity of the core, ARc is the axial rigidity of the columns and d/2 is the distance of the columns from the common neutral axis.
This concept, of the fully composite structure, will be used as a standard of efficiency when comparing different outrigger systems later in this paper. It is usually convenient. as well as necessary from strength considerations, to make the outrigger at least a full storey depth. Their bulk, and obstructive configuration, often makes it appropriate for them to share the plant room levels. This and aesthetic architectural considerations, as well as structural factors, will all have to be taken into account in arriving at the eventual number and location of the outriggers. For the structural engineer's contribution to the discussion it would be valuable if he could easily assess the relative performance of systems with different numbers of outriggers at different locations. He could then make known the structural and cost penalties incurred by departures from the structural optimum.
t
-l ~
i 1
The belt-hrnced structure hns been studied previously hy
2 d /2
AE M
Onyemaechi Nwaka graduated with an ~{.Eng. from McGill University in ·1977. This paper is based on the results of a research project which contributed to that degree. He now works in his home country, Nigeria.
...., .. "
1 ItA
In this paper, multi-outrigger structures are studied to obtain guidelines for the relative performance of different arrangements. Generalized formulae are developed for estimating the core moments, the column forces and the total drift in structures with any number of outriggers at any levels. General formulae are also derived which allow the optimum location of the outriggers to be determined. The formulae are used to give numerical. comparisons of efficiency for a range of structures with alternative arrangements· of single and multi-outrigger bracing. Some general conclusions are then 'drawn 'relating to the useful number and locations of outrigger bracing. It Rhould be restated that the Assumptions u5cd in developing the formulae in this paper, in particular the vertical uniformity of the structure, ~he flexural rigidity of the outr~ggers nnd the uniform distribution of the londing) re~tricts their use to the preliminary stages of design.
~ ",.
ThC' mnin vnl\1(' of tlH' work IH thcr('rore 1n thn p,cncrlll guldHIlCl! It pruvLul'H In numbcrlnR nnd locnt1.ng the oLltriggers for
their greatest effect.
'~
':~gt
"r-518 :
Stafford Smi.th and Nwaka
Multi-Outrigger Structures
519
j:
f.
ANALYSIS OF A SINGLE-BELT STRUCTURE
c) The outrigger action induceR only axial forces in the columns. The neglect of any bending in the columns due to rigidity o( their connections to the outriggers will be small and conservative.
The main purpose of this study is to develop general equations for the analysis of multi-belt structures. It is useful. however, to refer initially to a study of single-belt structures made by Tarannth (1). He developed equations for the outrigger restraining moment) drift and optimum location as follows.
d) The core is riRidly attached to the foundation. This will be ensured in the design o( the core to foundation connection. e) The sectional properties of the core and columns, and the distribution of horizontal loading) are uniform through the height. In a tall building this assumption will almost certainly be not valid. This reinforces the restriction that the results of this study are to be used for preliminary design guidance only.
Restraining moment (H 3 _ X3 )
w
M 1
;-;[ir + AE~d2J' wH
Drift
~
1
(H - Xl)
Analysis of Core Moments
M ~H2 _ X2)
4
= -BEl
1
1
2EI
in which w is the uniform horizontal distributed loading. the height of the structure and Xl is the distance of the rigger from the top of the building. The optimum location Xl of the outrigger for minimum drift is given by the solution of
4X~
+
3X~H
3 - U • 0
(4)
These equations will be used in conjunction with equations to be developed for two and three outrigger structures to establish general forms of the equations for restraining moments) drift and optimum outrigger locations. ANALYSIS OF A
DOUBLE-OUTRIGGE~
"
STRUCTURE
Assumptions These are The structure behaves linear elastically. This should be reasonably valid in both steel and reinforced concrete structures up to the design wind loading. In ~ases of reinforced concrete columns, a check should be made to assess the possibility of net tensile strt·sses in the columns. If thl8 occurH, nnypreoLctlolls for resisting moment and drift reduction based on ~ross sectional areas may be excessive.
~
"c~l.1l111
1I11.lInllll"
l,,·whlc·h I hc' oul rl)·aJ..'rli. "llh.""J.11 11"xllllllly
very Hllrf. :lrl' not rl",ld. Thfllr II .. xJhllltl(,:H will r .. d,u'p lit .. effectiveness o( the bracIng system. Thls important factor ls, t h l"' r ft f n r (', rl t Sf' \l H r; e cl R e pn rat e 1 v, 1 n t e r i nth e pnrc r .
Taking the restraining moments ~l and H2 as the redundant actions to be solved, the analysis can proceed using the conditions of rotational compatibility. Considering in Fig. 6a the rotation of the structure at the upper outrigger. the compatibility equation is H
e
a)
b) The outriggers are rigidly connected to the core and are thcmselveR flexurally rigi.d. Thi.R iR C1 groRR Rimplifi.cntinn of the
When a core stands as n free cantilever without outri~gerst it is statically determinate. The addition of a single outrigger system makes the structure once redundant. For each additional outrigger it becomes one degree more redundant. Referring now to the double cantilever structure shown in Fig. 6a) the structure is twice redundant. In addition to the "free l l bending moment distribution applied to the structure by the sideways force) Fig. 2b. each outrigger will impose on ~he core R restrAining moment which extends all the way froM that outrigger to the base, as shown in Figs. 6b and 6c.
1
2H l (H-X 1 )
-- +
c
Xl
2
(5)
w(HJ_X J )
HI (H-X l ) + H2 (H-X 2)
I
I
in which
'1
-
c
2) -12-.J: X (wx~ _ ~I )dx I _ -l.(",.~ 2 _ Ml-~12)dx"'() EI 2 I x E1 -
in which the first two terms represent the rotation of the upper outrigger and the last two represent the rotation of the core at the upper outrigger level. Equation 5 then reduces to
I
1.
£d 2 (H-X
I
2]
s= - + - - -2 TiT i\I\.c1 •
[
1
(6)
6EIS (7)
Considering now the rotation of the lower outrigger, it can he shown similarly that
i:~~
!'~!~I'I"" ~~ ':~f' j'
I
:1~
:..-.
Stafford Smith and Nwaka
520
3
+ M (l1-X )
M (H-X ) 1 2
2
2
MUlti-Uutrlgger
3
w
wH 6
(8)
6ElS
2
4
222
=--
(12 )
'3
M1
=
4X 1 ,+ 3X
(9a)
Xl
w 2 2 and (9b) M2 = 6EIS (H + HX 2 - X2X1 - Xl) \. Hl and H2 can b~ subtracted from the "free" bending moment diagram to give the iesulting bending moment distribution in the core, Fig. 6d. The forces in the columns just below the top outrigger are then! M1/d and, in the columns below the lower outrigger, ! (M1 + ~12)/d. The maximum moments in the outriggers will then be the product of the column force and the free length of the outrigger. Note that these will be less than the half moments Ml /2, M2/2 which are the moments of the column forces at the centroid of the core.
L[ IXl o
2
giving
6
wH
4
dx +
· x
I
2 (w~ 2-M ) .x.dx ' i
Xl
1
2
= 8El - ZEI [M l (H -
I
0
c
2 y 2 3 _ - 3X l X2 + 3•• " H - H - 0
( 13)
The procedures for deriving equations foi bending moments, drift and optimum locations have been given for a double-outrigger structure. Similar procedures can be used for structures with three or more outriggers to give equations corresponding to Eqn. 9 for the redundant moments, Eqn. 11 for drift and Eqn. 13 for optimum outrigger locations. For a triple-outrigger structure, these are Outrigger Moments
H
X
2
-wx
3
X - X 2 2
with respect to Xl'
ANALYSIS OF A TRIPLE-OUTRIGGER STRUCTURE
The moment area method is a simple means of determining the top deflection of the outrigger structure
= EI
3
2
1
~irst
The simultaneous solution of these gives the optimum values of Xl and X2 as O.3122H nnd O.6355H, respectively. These can he substituted into Eqns. 9 and 11 to determine the corresponding restraining moments ~nd minimum top deflection.
Analysis of Drift
tJ
222
l2(El) S
which, when partially differentiated, then to X2t yields
The simultaneous solution of Eqns. 6 and 8 leads to expressions for the redundant moments 2 2 6EIS (X 2 + X2X1 + Xl)
2
V_I
w[ (X +X X +X ) (H -X1)+(H +HX -X X -X ) (H -X )) 2 2 1 l 2 2 l 1 2 -2-
BEl
w
~lrU\;lUr~~
+
(w~ 2-~li-M2) .x.dx]
x2
(10)
2 2 2 Xl ) + MZ(H - X2 )J
(11)
-t
M)
Optimum Location of Outriggers The primary purpose of any bracing system is to resist drift. It is useful, therefore, to know at which levels the outriggers should he plRf:cO to nr.hirvp minimum rlri ft. T1H"HP """ he' fnul1c1 r Ir:al. In :
then to X2 . Therefore, substituting for M and H from Eq. 2 i into Eq. 11 gives
C)
w
= 6ElS
2
(X 2
+ X2Xi +
2
(14 a)
~~i)
__w_ 2 2 H2 - 6EIS (X 3 + XJ X2 - XZX l - Xl)
~. .,.
Hence, for a structure in which the dimensions, sectional properties and horizontal load intensity are specified, the actions in the structure can be derived from Eqns. 9, and the drift by substituting the obtained values of H and HZ into Eqn. 11. 1
by mlnlmll.lnJl, lilt, dC'lll'c:lloll f!'lunl Inn wlt.h r'·lIp.'c'l
Hi
w
2
= 6EIS (H + HX J
. (14b)
2
(14c)
- X1X2 - X2)
Drift
I
1
6
= wH
4
8J:.:l
_ _l_x [ l\{
2EI' 1
1
+
2 N (H 2- X )
2
2
+
2 2 N (H - X )
3
3
1
Optimum Outrigger Locations 'I'll,'"'' "rl' glvPIl hv XI' X'J. nnd X.\ frnm lhe' nnl"t'Inn nr l~qU:Il. IOIl~"
4X
~
(H 2_X2)
J 2 J + 1X X2 - X2 = 0 t t
( 15 )
522
~tanora ~mltn
ana
X3 _ 3X x2 + 3X 2X _ x3 1 1 2 233
3
22 X32 - 3X· 2X3 +03X 3 H
H
I~Wdl\.d
, l' I U • ". _. '"" .'-41 " . •
4X
o and the remaining
=
0
00""' · - .... - - - -- - --
32'3
1
~-l
+ 3X X2 - X 1 2
=
0
equations of the form
(16) ·
3
·22
Xj _ 1 - 3X j _ 1Xj + JXjX j +1
GENERALIZED EQUATIONS FOR MULTI-OUTRIGGER STRUCTURES Equations are now available to solve the redundant moments 2, g and l4}, drift (3, 11 nnd 15) and optimum outrigger locations (4, 11 and 16) in single, double and triple-outrigger structu~es. An inspection of these equations indicates sequences which allow corresponding generalized equations to be written for structures with any number of outriggers, as follows.
3
X'+l J
o
(20)
where j is equation number.
(Eqns.
Outrigger Moments
1
In a structure with N outriggers, the restraining moment due to the uppermost outrigger is given by
In the equation for a single-outrigger structlJre and in the for an N outrigger structure X2 = Hand XN+1 = H, respectlvely.
~th equ~tion
The formulation and simultaneous solution of the set of equations for a structure with N outriggers will determine their optimum locations for minimum drift.
THE
PE~FOR}~NCE
OF OUTRIGGER STRUCTURES
:1
I:
l l;. ~
"~
2 2 6ElS (X 2 + X2X1 + Xl) w
If the outrigger is at the top of the structure then Xl = O.
In an attempt to develop a feeling for the relative performance of optimum and certain non-optimum outrigger structures, the derived equations will be used to ohtain valueR of optimum locations. restraining moments and drift reductions.
In all other outriggers, i to N
Structures with Optimum Location of Outriggers
M1
=
(17)
w 2 Mi = 6ElS (X i +l + Xi+lX i - XiX i _ l ~+l
taking, for the lowest outrigger,
I
2
x.1- 1)
(18)
j
= H.
,I~
iI~I'ro 'tl.l ~~: ~ .i,
'f;· '~t
01 ~,~
:j." The results for optimum locations) restraining moments and drift reduction for structures with up to four levels of outriggers are given in Table 1. The presentation and implications of the results are discussed below.
:t,,·
l,:~
"i':·rl t.~
"
Drift The total drift for a structure with N outriggers is given
~
by
Ii
wH 4 BEl
___ _
1
N
r
2EI ~i=l
M (H 2
· 1
2
.(19)
- Xi )
The first term represents the drift of a free cantilever under the action of a distributed horizont~l load wand the second represents the reduction in drift due to the o\ltriA~cr syRtem. Optimum Location of Outriggers '1' II U
U
II L J IIIlJ III LtJ t: II L LUll tJ. X1
llJ
XN'
t)'
l
II C
(I 1I
l
r f ~ ~ l~ .. Ii
J It
•, "
N
1
Outrigger locations Values which are obtained from the simultaneous solution of sets of equations. as per Eq. 20, are optimized to produce maximum drift reduction. Although the drift reduction will also be influenced by the axial stiffness of the column system -the larger their stiffness the greater the reduction - it is evident from Eq. 20 that the optimum locations are independent of both the core and column stiffness properties. The resulting optimum locations are noteworthy in two respects. First. that the top of the structure is not nn optimum locntion for an outrigger in any of the cnsey. Second, thnt in nll cases the optimum locations are very close to the equal interval points in the height of the structure. For example, the optimum level of the outrigger in n HinKle outrigger structure is almost exactly at mid-height whilst those in a three-outrigger system nrc cl o~e to the quarter, hal f an1d three-quarter heights. Although th~ primary function of l1 bracing system is the reduction of drift, an associated merit of the o\ltrlp.Rcr bracing Rystem is the reduction in core moment. I': q. 2. r() r 1\ H In ~\1 C 0 \l t r 1RRc r H t r \l C t \1 r c. ah ow 8 t hn t the low crt he 1{t!9training moments
outrigger structure can be obtained from the simultaneous solution of n set of N equations including a first equation.
J1
.. :~
il~I'n':11 ••
~~I
:
IW:·l ,!\
:1
I~;:
q;~ 1i;:~
j·:X
11'~1
ii:
:!: "
~~lLll
.\il~,~ IIJ~ . !;~
!
I'
d 'I
.- ...
,---
.- .Stafford Smith and Nwaka
'-524
Multi::Outrigger Structures
outrigger the greater the resisting moment it provides. The benefit of placing it very low on the structure to give a high resisting moment is offset by the fact that it would not be as effective in reducing qrift nor would the reduction in moment, from the outrigger to the base, extend over much of the height of the core. Indeed the moment in the core above a very low outrigger might even be larger than below. The values of resisting moment in Table 1 were obtained from Eq. 17 and 18, and correspond to outriggers at the optimum drift locations. The resisting moments are expressed non-dimensionally as a percentage of the fully composite moment, Mfc (Eq. 1) .. The values are, therefore, absolute measures of the moment resisting efficiency of the system, independent of the column and core sectional properties. The justification for this form of pres entation is as follows 2 AE d /2
from Eq. 1 1-1 f
c
•
EI + AE d 2 /2 c
Mfc
= EIS
I
1
The results in Table 1 show that the larger the number of outriggers, the greater the'total moment reductton in the core. However, for each additional outrigger there is a diminishing return in the extra reduction of moment. The largest resisting moment of an individual outrigger occurs in a single-outrigger structure. In multi-outrigger structures the lowest outrigger carries the highest resisting moment and each successively higher outrigger carries less. Reductions in drift These are also given in Table 1 in a non-dimensional form, as a percentage of the drift reduction which would be achieved by the same system acting fully compositely. By simllar reasoning to that for the restraining moments, the fully composite reduction in moment due to the outriggers and columns would be
i
2 ( 21)
2
2 wH · '-2-
1
or
wH
c
j
?
AE d-/2
J
dLi
fc
c
• -
wH 2
EI + AE'd /2 c
wH
(22)
4 ( 26)
BEl
4 (27)
8(EI)2 S Taking Mi' the actual resisting moment of outrigger i from Eq. 18, and expressing it as a percentage of Mfc gives
*
Mi
Mi
= M:
fc
~ 100
100 w
2
= 6ElS (X i+l
2
+ Xi+1X i - XiX i _ l - Xi - l )
2EIS
·----2 wH
(23)
hence
*i =
JJ.J
M
(X~+l
+ Xi+1X i - XiX i _ 1 2 H
X~~l)
(24)
I i
Taking the second term of Eq. 19 as the actual reduction in drift, d6, and substituting in it for Mi from Eqs. 17 and 18. it can be expressed as a percentage of the fully composite reduction
by
~
d/\
Mt
'M~
Mi
= 100
•
wH2 ET~
where S is defined by Eq. 7.
dl\
'If
di\rt'
:< 100
~ 1(
Therefore, although the actUAl resisting moments of the outriggers are dependent on the size and spacing d, of the columns, the efficiency of the system, based on the fully composite hehnviour of the columns and core as heine 100% efficient, is ShO~l hy Eq. 24 to be dependent only on the number and locations of the outriggers. Values of in Table 1 can be converted to Actual moments ~11 in n pnr t iculnr struc t ure hy us inA
525
dl\
I
(:? '))
1 ':'
N + E 2
8fEI)2S wH 4
w
100 x ~~-' ·
(X~+l
[(X
l2(El) S
2 2
2 2 + X2X1 + Xl) (H
+ Xi+1X i - XiX i _1 - X:_ 1){H .
r'" * - 66.7 --,--II
2
2·
2
r (X " + X" X, + XI) (11
I
•
2
•
'2
- XtX i _1 - Xi_1)(H
2
- Xi)]
2
2
2
- X I)
-
X~)]
2 X) 1
(28)
~ 2 + f. (X, 1 1 + X 1 I. \ X1 :~
(29)
:>..-rr-~
-
.:)IdTTOfO :7IT1lln ano
-
J~Kd--
0-
O_---J~/
Generally, outrigger flexibility will shift the optimum locations of the outriggers, however many, up the structure from the posit ion s Cl u0 ted for r -i g i d 0 U t rig gerR. The ext en t 0 f the shift will depend on the rcductlon in stlffncHs or th~ outri~n~~r and also on the stiffness of the column system. The more flexibl~ the outriggers, the greater the shift. On the other hand, the stiffer the column system the greater will be the effect of any change in outriAgcr flexihility.
Equation 29 demonstrates that the efficiency of the system in reducing drift, taking the .fully composite system as 100% efficient, is dependent only on.the number and location of the outriggers. The efficiency~of the outrigger system in reducing drift is so high, 88%foro~e outrigger and 96% for two, that the extra return for additional outriggers diminishes even more than 1n the case of restraining moments. The extremely high, 98.6%, efficiency of a four-outri~gcr system implies that nny more than four is not worthwhile.
As an illustration of the influence of outrigger flexibility, Table 5 shows the change in the resisting moment and drift reduction efficiencies for structures with three values of core··to . . outrigger stiffness ratios, and three core-to·column stiffness ratios.~. Clearly,the influence of softening the outrigger is greater for structures with the high column stiffness Q = 1 than with the low-column stiffness, C1 = 10.0. Fig. 11 shows the shift in the optimum location of a single outrigger as it becomes more flexible. for hypothetical structures with axially 'rigid' columns. Two points are plotted in Fig. 11 for existing structures with known core-to-outrigger stiffness ratios, which show that even in the extreme case of assumed rigid columns, the shift in the optimum positions for these structures is not large.
Structures with Outriggers at Even Spacing
a,
To allow integration of the outriggers within the normal floor intervals. as well as for architec~ura1 and plant considerations,it may be more convenient to lo~~te the outriggers at equally spaced positions rather than at the optimum locations. Therefore Tables 2. 3 and 4 present the resisting moment and drift reduction efficiencies for equi-spaced outriggers. It is useful to compare their performances with the optimum. Table 2 compares the efficiencies of systems with a single outrigger located either at the top, or at mid-height, orat the optimum location, Xl = 0.455 H. Whilst the mid-height and optimum outriggers are Jndistinguishable in performance, a single, top outrigger behaves relatively poorly in providing only 60% of the moment resistance and 75% of the drift reduction of the same outrigger with an optimum location. Figures 7 and 8 compare the bending moment diagrams and deflected shape for top and mid-height single outrigger structures with those fqr a free cantilever. The core to column stiffness ratio a (where a = 2EI/ oAE c d 2) for this case is 1.0. The fully c~mposite, that is maximum possible, reduction which could have been achieved in these cases for moment and deflection is 0.5. Tables 3 and 4 are for· equi-spaced, multi-outrigger structures, the first including one at the top and the second without. It is evident that for systems with the same number of equi-5paced outriggers,those with an outrigger at the top are relatively inefficient. Indeed, the total resisting moment of any system which includes a top outrigger can b~ almost equalled by a system with one less outrigger, omitting the one at the top. Figures 9 and 10 compare the bending m~ment and deflection diagrams for a triple-belt structure with those for a free cnntllever core.
CONCLUSIO~S
The followinggeneial conclusions apply:
Allthc dcrlvutlons and results have safar assumed the outriggers to beflexurally rigid. In reality, their flexibility will reduce the overr.111 Rtiffnp.RR of thf' n\ltrtRRl"r HyHI.(\1ll /llul
J.I
Outrigger bracing, in single or multi-outrigger arrangements, is an efficient means of reducing drift and core bendinR moments in tall building structures.
2.
In single-outrigger structures, it is more efficient to locate the outrigger in the mid-height region than at the top of the structure.
3.
Multi-outrigger structures can be more effective than singl~ outrigger systcmsj however with each Cldditional olltrigger the increment of additional stiffness and moment reduction diminishes. The possible reductions in drift for more than four outriggers are insignificant.
4.
If, in multi-outrigger structures, the outri~gers ore located
I
~:
fa t (! q \I I cl {H t nn t III Lg h t {11 t (' r v f\ 1H r r n III t 11 e J~ r 0 lJ n d. h 1I t nmit t 1n g une Ht the top of lhu structure, u close to optimum rcuuctlun in drift will be achieved. 1
'1
An investigation is proceeding into the detailed effacts of outrigger flexibility, but some general comments based on preliminary results ~an he mArtp..
1. !
INFLUENCE OF OUTRIGGER BENDING STIFFNESS
'."11ke it lell see fee tive in controlling the d r H t and core moment II
NlOl(f.UUtrrggeT-~1.rUClU-r~~
I
~~
Multi-Outrigger Structures
Stafford Smith and Nwaka
528
529
REFERENCES
1.
~
Taranath, B.S., "Optimum Belt Truss Location for High-Rise Structures", Structural Engineer, No. 8,V53, August, 1975.
c
o
or+oJ
U
U
'f-
~<)
2.
3.
McNabb, J.~., and Muvdi, B.B., "Drift Reduction Factors for Belted High-Rise Structures", Engineering Journal, American Institute of Steel Construction, Third quarter, 1975. McNabb, J.W., and Muvdi, B.B., Discussion of Reference (2) Engineering Journal, American Institute of Steel Construction First quarter, 1977.
"C"O
I
C1I
J,.'f+J
area of column horizontal distance between columns
d
a:-
\C
ID
.......
en
\G'i
~ f'.: .......
0)
0'\,
VI
x
CIt,
0i: '" Oft'
I
I
~
T
h:
~
'7
Ln ~
\0
It
:F
.....
~
::')
fU
o
:F
c:::
Vl UJ
'0:
.. rr:
...
t··
,Jrdl
N
;1-'1'"
intensity of uniformly distributed horizontal loading
~
~
co
(\.
0IIl:t
of[
\0
M
f
::')
Q)
....
0:
-4: r-4
VI
0:
LLJ
~
~
C' C
distance of outrigger from top of structure
c.!:l
..
"Z
C?
f',
:;'
'"
Ui)
-:e.-
,.rJ...
N
~
o
co
U).
I:::J
0')
oI
V\
E c
elastic modulus of columns
H
total height of structure
I
moment of inertia of core
ex
M fc
resisting moment of outrigger i
Mi expressed as a percentage
N
s
a parameter
VI + AE:d 2]
~ ..,..
.-:r ..
LIt
~
~
-
horizontal drift at top of building
d6
reduction in drift
0-
• v
~J ~
6
J
.
Cully composite reduction in drift
I
centage of dli fc
":'!:
~
l.L.
::::
0: Lt.J
0..
~g
\U
o
~ ~
Ln
a
M 4l:t' IIIl:1"
o
t---.
~
o
U'l ll;:I"
Q.
N
t'""
('\,
M
~.
C""!
o
C')
~ Q
C'.:
r",
q-
4'
C" L 0' GJo,... .J:J'-
t: ...,
;=?
R
Lt.J
...J
co
.... U')
,II
I
c(
V\
d6 expressed as n per-
.... 0..
o
it- L
d6*
:~l~-
~
'
g ,.,:.' ~
o
o
C"'-
C":~'
~ ~"i.I
i·i
1/
.. \
l.L.
+oJ
l5 L
I
w
W
'+-
i~
o
:::t:
5 1+;-
"'~I ~'1
";1'
•• ~ 1:1:
;\;1
~ ~
it )of(-
dL\fc
(;)
"i
;~~r , .\
.......J :::J 2::
'
V\ .,...
of H
fc total number of outriggers
core to outrigger relative stiffness parameter EI . d (EI)o H
=-
fully composite resisting moment 6
N*
core to column relative stiffness parameter = 2EI 2 AE d c
~
B
i
c-
:1
::, ;~~.
0:
.~
GI
N
Q
eo
co
o
elastic modulus of core
E
0-
'to- ,;-
HOTATION A
o
~l'
:·t
III I
'/
..
~
CJ ~
c:
,
Resisting moment as % of Mfc
location of outriggers
Drift reduction as % of d~fc
d~ (J ,...
~
Optimum X1/H = 0.4554
55.4
...=
o
88.0
c
CI
3
~1id-Ueight
58.3
X1/H = 0.5000
~
::J
87.5
~
::s
c
Top-Hat X1/H ::: fJ
..
~
f
I t
'!umbp.r of C'.Itriq
--
* Xl
, •
X*z
~
~ ~ ~
~
Distance from toP :
I
:2
66.6
PERFORMA~iCE OF SINGLE-OUTRIGGER STRUCTURES
TABLE 2
....
33.2
! ~
*
:<3
....1------ ..
Po
I i
*
XII
r
•
R~sistinn
\1 *
1
:
·.1?*
.--~::.;,:=,'=~'"
. ..
noment as
*
"'1
~
till·
...
.
of Mfc
'*
f
;
Mil
I.~~
Drift reduction dd as .~ of dt\fc
::;
I
i I
f
I
1
~
-
-
-
33.2-
2
()
0.5
-
-
8·3
t
..
.~
-
-
JJ.2
66.6
-
-
5S.,"3
91.6
oc ,.....
!
,i I t
50.0
:!.
QQ
I
(JQ
I
1 I
,
3
0
0.33
0.67
4
~
.Q.25
f).
-
8.6
23.1
4-3.9
-
70.6
96.3
0.75
2.1
12.. c;
2.5.0
,7.5
77.1
97.8
I
50
..
_--J
TABl E 3
PERFORMA:JCE. OF EQU I- SPACE') I)UTR I ~GC:R ST~UCTURES ~'IITH TOP 11UTR IGfiER
... r.r. ,.... ...c. !1 c... m to
I
iI
~
C
(I)
o
u.
~fflit£tiii'iii-_fl.I1_~d4t
• •Ii-WEg~~¥~::·:~·······
c.n
~I • I
f
Distancp.
'tumber of outriggers
i
Xl*
top : H
X,*
X*2
-
0.5
1
fro~
Resistino X* 4
t.4
I
'1
-.,!*
as
~
*
I
&.1 * 4
:
'~3
;
~fc
of
1:"1*
Drift reduction d6 as ~ of d6 fc
til ,..... ~
-
-
*
mo~nt
5S.3
-
-
-
87.5
5~.3
~
o
-t
C. 2
0:33
0.67
-
-
2'5.9
~4.
-
-
70.3
95.5
3
0.25
0.50
0.75
-
14.~
25.0
37.5
-
77.1
97.8
4
0.20
~.4
0.6
0.9
9.3
16.0
24.!)
32.0
81.3
98.5
.,,1
til
3 ;::;:
::r ~
::J C.
Z
~
,
. TABLE 4
}.. - - -
-t ,
PERFOR~A!lCE
10
EXCLUOI~G
.:.-
~
;::-
TOP OUTRIGGER
~
.-......--
.'.~'='.~
- ._.. - -._ .. _-- .. - ..... --------------------_.---Drift reduction d6 as % of d6 fc Resisting moment as % of Mfc ...
....
o
..- - _.._ - - - . .
hr·
.....
N
OF EQUI-SPACED OUTRIGGER STRUCTURES
. ..- ..... -..
1.0
I I
2.0
1.0
10.0
2.0
10.0
66.6
....~ oc•
33.2
33.2
33.2
66.6
66.6
32.0
32.4
33.0
64.0
64.8
66.0
51.9
61.6
23.6
a
C
~
,..... :!.
26.1
47.0
30.8
= Core: column stiffness ratio
O"Q CJQ
....
(D
,.....
CIl
2 AE d
-I
~
t: ~
2EI
C
-I
B = Core:
outr;~qer stiffness ratio
E1
• d
{fI}o
IT
(t)
en
TABLE 5 INFLUE~CE OF SI~GLE OUTRIGGER FLEXIBILITY O~ PERFORMANCE (JJ
W VJ
CJ1
-+
OUTRIGGERS
~.
--...
I
-.
HAT GIRDER
I
-+
c.n ,... ~
COLUMNS
---h. ......
o ., c.
8.M.
,', . :i:
II
I
I
Fig. 2a--"Free"
Fig. 2b--B.M .
cantilever
diagram for a
_.
3,...,.
free cantilever
CORE
I
CIl
:r ~
:::J
C.
Z ~
~
~ ~
Fig. 2d--B.M. diagram for outrigger braced
Fig.2c--Outrigger braced cantilever
Fig. I--Jutrigger braced structure
cantilever 1..---
__..
~.,
1201t
fO 1 404
I
41~
I
•
•
z..-::'l...r:·......·
_'" ,.
tJ. *",lSj';O_.,.
,r.
OUTRIGGERS 2 - STOREYS
..
..
OUTRtGGER BENTS
DEEP
r+ A
FLOOR
....
+-
o (\J
~
c 2! •
•
+-
oc
'f-
o
...
.....
o
~
......
:::!•
eJ r
C1Q
()Q
<.0
4A 5
e
40 ft
11:
200 ft
I
... ,.... ...enc Q. c...
to
tt>
U')
Fig. 3b--Typical plan of first Wisconsin Centre Fig. 3a--Section AA of first Wisconsin Centre
~:.~;}~:~~Y:---=
(J1
~
. ~__·:.}/:;:~~::;~~~1!!~ri-~:.:
- - - - -
-
. - '.'Stafford Smith and Nwaka
536
HAT GIRD
---
,.-
43 rd FLOOR
;,0
-. OLUMNS
.
4
_.... _21· ..
""V~
...
'. L....:.-
I
I
~
'" ~
~-
.,
I A
L
--+--+-
,
I
1 --,
\
I
11
8
--+ --+ ~
...!L
.: "~
537
CORE
35
BELT GIROE
.- - Multi-Outrigger - - Structures
-
INFINITE NUMBER OF OUTRIGGERS
.A.C. WALLS
}~.,~=
FOUNDATION
~
;
-.
,(
A
.
•
Fig. 5a--Fully composite outrigger braced structure
Fig. 4a--Typ;cal section of Yasuda Building
FREE CANTILEVER
B. M. DIAGRAM
OUTRIGGERS ....,..----.
:/
I
~
E
;t
~
J
• • ~~l
I~
E Q
. ~"l~_' - -_ _~1_m_.
Fig. 4b--Plon of Y4sudo Uu11d1ny
.-,ty..m~
B
B.M. IN CORE B.M. CARRIED BY COLUMNS
B.M.
FlU. ~Jb--J\11()CilLI()1l of hendlnu lI1o\li(!nl 'n fully composlLe structure
JI
':',
":\J~\t~!\: :,1~!11 ' .
j,",t'.:
."
..
~
_
,._ . . . . . . . . . v .
~~-
-'~-
.~, ~~.~~~~~~~;~~:~:~~~~;~~~ _:·:Z-:i... _•. _-- ...._ ... ~.~:.:: .-:_":w:.:• •-..·.,..~ . '::~7~: ..
l
en ."
r t .-.
-.os co c· n
!
~.
~
~-. _~.~:::::~:~.-:-:-=-' ._~.: --.;~_:
~r 20
c-t" 0'\
c:s:u -.os.
-t ~
I Z
~
o I o
fllGl
C
(i)~
t+
-
J.
..
I
-1-0
(,Q (,Q
\0
rn
CD
-s
en c..
•• :::c (1)
-'" c
r-+
c:» ::s
~
CJ
3:
",,.~ ,." NI ~
I
X
L
:r:
:0
·le
N
:s:
~
...0..
o
en
~
r+ ....... ."
o:::::s ....... ....... u:l co ::J3 0...0 m
.
3:
-30"
-s c:»
....... :::::s
-
~
:r
\0
Q..
3:
:::s n
......
:=:
S
3
CIl
3
I
N
I
:::c:»
L
NI~N
NI~N
~
CO
-l> -i o
CDI
-r'1
(JJ ~
C PI
.~
c+::J
a.
o
Z
~
PI
0...
~
~
.- ....... m
N
:J C.
010
Ul I C-SI
0
'"1
r+CI::O
(1)
-s3C'D
....... c.n c.n \0 r+
co
c...-.os
-.os
CD-
3: N
C'DCPI VI
::s
I
I
0..-_..
••
\
~.- ------6
2
tat
_
TOP ,
\
-rR..
TOP
0.1661---0 \
..
O. Loading w/tnt : 0.1 ~ Height _
0.1
\0.2 \
C
0.2L
~.3
O5 '
0.3
FREE CANTLEVER
\
A
~
C
0.4
:x:
::x: +
0..
0
t-
~j ~
0 0:
LL.
OQ
0.8
--'I
cr. ,.... --t
C
A
0.8.
a
0.9
0.9
BASE 1.0
CD
0.7
Ij
0.7
(f)
I
o
,
,
02
,
!
0.4
t
!
0.6
>
A.
0.8
!
>t
LO
2
BENDING ·MOMENT IN C.ORE + wH /2
Fig. 7--Bending moment diagram for single-outrigger structure
:~~~;;,.:.;~:::-:-.
--'I QQ.
0.6
~
~
o
E
0.5
+
SINGLE OUTRIGGER MID-HEIGHT
0-
g
S oc• ,.....
BASE 1.0
c--.r
2EI
d..
to
AE d 2 = 1
(I)
c
I
o
I
!
02
I
I
0.4
!
!
0.6
!
•
,
08
I
LO
4
HOOZONTAL DISPLACEMENT + wH /8 El
Fig. 8--Deflection diagram for single-outrigger structure
CJ V.
c.c
.~~~~~il.I~~~:5='~Jjf~::'~~~'~i~~;~;~::~~
. ...._.:-::. . _".- _.:.. ~
~ _.-.-~'~. ~~~
~
-=_r~" •
..
.
~ ~
:~.,j.:,I~~"''''' .,,,
~
~
•
..
>'. ••
#
._
-,...
'"
.....
.r , -.
•
.. •
. #
'"
~~~~~~~~~~~~~~0~,~~~~~~~~~~~~~~~~~~h
TOP
0
TOP
I
2EI
do. = 0.1
\
TOP OUTRIGGER
8.M. DIAGRAM FOR
\
0.4 ..
0.6
~u
07
'"
,, ,
g
~-~ ,
if.X
3
(f)o-
'" 0.9',
I
!
o
,
Q2
,
~ 0.75 H
,
"" , ,
-
,
0.6
04
,
,
08
'"
I
BASE 1.0
LO 2
AE d 2
I
,
,
,
I
1
I
I
~
= ,
,
I
o 0.2 04 0.6 08 LO HORIZONTAL DISPLACEMENT e;. wH4;e EI
Fig. lO--Deflection diagram for triple-outrigger structure
for triple-outrigger structure
0.
d-.
c
Fig. 9--Bending moment diagram
......
2EI
I
BENDING MOMENT + w:H /2
....
&.1 ~
0.9
: "
.=1
HeicJht ::.
BOTTOM OUTRIGGER ...
11°·8
CANT1LE~lR
-
U>oding w IUit
0.7
:r Z
aM. DIAGRAM FREE
"
0.6
o
fE
::;:
FREE CANTILEVER
~
sanOM OUTRIGGER
0.8',
BASE 1.0
31
0.5
a..
~
~
-t
c. +
"
t·
I
....
::I: '
...
•
..
.. I~
n -n
0-
--'\0
c: •
5.Con
0 tv 0
.-
••
0
-0
C'+
3 c: 3
0
::r ro
0 W (J1
0
0
0
(J1
~.
h -
.I
YI!;.. ()
-.
\0
~
--i ~o C ::0 ..
::r
C
» -~ r G> (J1
C'+
0 -t)
--•
::;-
0
O(Jl
-t)
~O
-..II
~
--
).
CT
--"
CD
C ,...,.
::Oc C-f 0::0
C1Q
CG)
tn
::!.
(JQ
-f-
0
c
rt
-e til -e
::oG)
:l
,..
fTlfT1
to to CD
Ul~
t:
""1
--.-
n
0 ""1
1"'+
-.
t» X
~
--..
'<
-s
0:> 0
c:
GZ1
~
t'D
en
t .O
a-
~U
0 0
,,:',::::'.,-
I
I g:~ I
0.4
\
~
~
0.3
X2 t X3
MIDDLE OUTRIGGER X2 - O.SH
,.__ .1
+ 0.. o....
\ \
0.5
YII :. x·: It:
-_-
TRIPLE OUTRIGGER STRUCTURE -.~.,
\ \
lit
w J1Jnit Height
~\
\\
O.2~ ~m
r· Loodi"9
0.25 H
~
° 0.1
c
XI: .J ~
:I:
AE d 2 -
a
02"
0.3
_
I
CJ1
~ -4
d:-c!-=~·:'~:~:': '~~;";'~:~7~~::~~"'~]:;-' ;:;~.;;'.i~~~~~~~~~~.~::~~~~:~?-'~4-S't~~!!fii/fr~~-::t~?~~!.4#f3:~~~~f!l~~~~¥f:::'-;~"~
I
9 Shear lag effects in buildings 9.1 The shear lag phenomenon Shear lag is the lagging behind of bending stresses relative to that induced by pure bending action (as derived by the plane section remain plane assumption). It occurs whenever there is shear stress and is generally more significant for structures with relatively small height/width or length/width ratios. Shear lag leads to stress concentrations around the comers of the structures and additional deflection. Asa result, the overal structural efficiency of the structure would be reduced. Common structures whose structural efficiencies are affected by shear lag include framed-tube structures, shear/core wall structures and bridge deck structures. Shear lag occurs· in both the web and flange panels, but since the shear lag phenomenon is generally more significant in the flange panels, most researchers take into account only the shear lag effects in the flange.
41
9.2 Shear lag in framed-tube structures See attached paper.
9.3 Shear lag in shear/core wall structures See attached paper.
42
SHEAR LAG IN SHEAR/CORE WALLS
By A. K. H. Kwan l ABSTRACT: Shear lag occurs not only in bridge decks and framed tubes, but also in shear/core walls. However, there have been relatively few studies on shear lag in wall structures. Moreover, most existing theories neglect shear lag in the webs and, although they are acceptable for bridge decks that normally have flanges wider than webs, they may not be applicable to shear/core walls whose webs can be much wider than flanges. To study the shear lag phenomenon in wall structures, a parametric study using finite-element analysis is carried out. Unlike previous studies that neglected shear lag in the webs, many layers of elements are used for both the webs and flanges so that shear lag in the webs can also be taken into account The results indicate that the shape of the longitudinal stress distribution in an individual web or flange panel is quite independent of the dimensions of the other panels. Based on this observation, design charts and empirical formulas for estimating the shear lag effects are developed for practical applications.
INTRODUCTION
Shear Lag Phenomenon The Bernoulli-Euler assumption that plane sections remain plane after bending is often used for the analysis of beam structures. According to this assumption, the longitudinal stresses in the webs and flanges should be linearly and uniformly distributed, as in Fig. 1(a). However, this assumption is approximate and strictly applicable only when there is no shear force or when the structure has infinite shear stiffness. In actuality, when the structure is subjected to shear forces, a shear flow would be developed between the web and flange panels and, owing to shear deformations of the panels, the longitudinal displacements in the parts of the webs and flanges remote from the web-flange junctions would lag behind those at the junctions. As a result, the longitudinal stresses in the webs and flanges would become distributed as shown in Fig. l(b). Such "shear lag" phenomenon reduces the effectiveness of the webs and flanges and may significantly increase the longitudinal stresses at the web-flange junctions and the lateral deflections of the structure. Shear lag is most pronounced in beam structures with relatively wide webs and/or flanges such as bridge. decks and shear/core walls, and in beam structures with low shear stiffness such as framed tubes. However, although there have been a number of studies on the shear lag phenomenon in bridge decks and framed tubes, there were relatively few studies on shear lag in shear/core walls.
Existing Analysis Methods Existing methods for shear lag· analysis include (1) the folded-plate method; (2) the harmonic analysis method; (3) the finite stringer method; (4). the finite-element method; and (5) semiempirical methods. In the folded-plate method (DeFries-Skene and Scordelis 1964; Kristek 1979, 1983; Kristek and Skaloud 1991), the structure is treated as an assembly of plates interconnected at their longitudinal joints. The displacements and forces along the longitudinal joints are expressed as Fourier series of harmonic functions and by considering the plate bending and membrane actions of each individual plate, each tenn in the series for the joint forces is related to the corresponding term 1 Sr. Lect., Dept. of Civ.. and Struct. Engrg.y Univ. of Hong Kong, Pokfulum Rd., Hong Kong. Note. Associate Editor: ~c M. Lui. Discussion open until February -1, 1997. To extend· the clOSing date one month, a written request must be filed with the ASCE M~ager of Journals. The manuscript for this paper was submitted. ~or review and possible publication on November 22, 1994. This paper IS partor the Journal of Structural Engineering, Vol. 122, No.9, September, 1996. ~ASCE, ISSN 0733-9445/96/00091097-1104/$4.00 + $.50 per page. Paper No. 9648.
in the series for the joint. displacements by a stiffness matrix. Then, by also representing the external loads in the form of Fourier series, the entire analysis can be conducted separately for each term of the series and the final results can be obtained by summing the partial results. As for the folded-plate method, the hannonic analysis method (Abdel-Sayed 1969; Kristek 1983; Song and Scordelis 1990a,b; Kristek and Skaloud 1991) also represents the external loads as Fourier series. Unlike the former method, however, this method simplifies the analysis by. neglecting the outof-plane bending action of the individual plates and treating the web plates as simple bending elements so that the analysis can' be confined to the flange plates only. The finite-stringer method (Evans and Taherian 1977, 1980; Taherian and Evans 1977; Connor and Pouangare 1991) also treats the web plates as simple bending elements so that the analysis can be confined to the flange plates. Instead of using Fourier series solution method, it models the axial action of the flange plates by a finite number of stringers welded onto the plates, and the shearing action of the flange plates by the plates themselves that are assumed to take no axial loads. Thus, the axial and shearing actions of the flange plates are separated and the governing equations become easier to solve. Using only three stringers to model a flange plate, a simplified version, called the three-bar method, has been developed by Evans and Taherian (1977, 1980) for practical applications. The finite-element method, being the most powerful and versatile numerical method, can also be used to evaluate the shear-lag effects. Moffatt and Dowling (1975) had, by using the finite-element method, carried out a comprehensive para-
flange
web
Ca)
flange
web
(b)
FIG. 1. Axial Stress Distribution in Beam Structure: (a) with No Shear Lag; (b) with Shear Lag JOURNAL OF STRUCTURAL ENGINEER1NG./ SEPTEMBER 1996/1097
~,.-
metric study on shear lag in bridge decks.. It was found necessary to use fine mesh divisions over the width and length of the flange plates.. However, the web plates were assumed to behave in accordance with the elementary theory of ben~ing and, thus, each web plate was modeled by one layer of elements only. From the finite-element results, Moffatt and Dowling (1975) had produced a set of design values for the estimation of shear lag in bridge deck structures.. Apart from the previous method, there are also some semiempirical methods based on energy formulation (Coull and Bose 1975; Coull and Abu EI Magd 1980; Kwan 1994).. In these methods, various simplifying assumptions regarding the longitudinal stress distributions in the web and flange plates are made to render the analysis more tractable, and solutions are effected by minimizing either the strain energy or total potential energy of the structure.. They vary in their accuracy and are generally not as accurate as the rigorous methods.. Although shear lag can be accurately analyzed by many of the existing methods, such accurate analysis is generally very time consuming.. For practical applications, there is still the need for simple methods, which can allow quick estimation of the shear-lag effects without the use of computers, particularly during the preliminary design stage. There are, however, very few simplified methods for shear/core wall structures. Moreover, in most of the existing methods, the webs are assumed to act as simple bending elements and, as a result, any possible shear lag in the webs is effectively neglected. Although methods that neglected shear lag in webs are acceptable for bridge decks that nonnally have flanges wider than webs, they may not be applicable to shear/core walls whose webs can be much wider than flanges. It is, therefore, necessary to have a new method that is easy to use and yet capable of taking into account shear lag in both the webs and flanges for application to wall structures.
Present Study Herein, a parametric study on the shear-lag phenomenon in shear/core wall structures is carried out by using finite-element analysis. Unlike previous studies that neglected shear lag in the webs, many layers of elements are used for both the webs and flanges so that any shear lag in the webs can also be taken into account. Apart from point loads and unifonnly distributed loads, triangularly distributed loads, which are quite common in tall building structures, are also considered. The ultimate aim is to develop a simple method for estimating the effects of shear lag in both the webs and flanges of shear/core wall structures.
PARAMETRIC STUDY USING FINITE ELEMENT ANALYSIS
Finite-Element Analysis To study the shear lag phenomenon in wall structures, a parametric study is carried out by analyzing a number of core wall models using the finite-element method. The models are shown in Fig.. 2(a). Due to symmetry, only half of the core wall model is analyzed. Very fine element meshes of isoparametric eight-noded quadratic serendipity elements are used for both the web and flange panels, as illustrated in Fig. 2(b). Twenty layers of elements are used in each web or flange panel (since only half of each flange is included in the analysis for taking advantage of symmetry, 10 layers of elem~nts are used in each half flange). Along the height, there are 60 elements in each layer which are distributed in such a way that 40 of them are evenly distributed within the lower half while the other 20 are distributed within the upper half. The structural parameters studied are the flange width/web width ratio, the t
1098/ JOURNAL OF STRUCTURAL ENGINEERING / SEPTEMBER 1996
height/web width ratio, and the height/flange width ratio.. Altogether, 15 models are analyzed.. They are numbered from 1 to 15, and their (web width:flange width:height) ratios are given in the second column of Tables 1-3.. The flange width! web width ratios of the models range from 0.. 33 to 3.. 0, while the height/half web width and height/half flange width ratios both range from 3.. 33 to 40.. 0 . On the other hand, the web and flange panels are assumed to have the same and constant thickness along the height of the structure, and the Poisson ratio is taken to be 0.. 25 throughout. Three loading cases, namely point load at top, uniformly distributed load, and triangularly distributed load, are considered. The loads are applied laterally
20 ale.ents 1n upper
halt H
40 elements 1n lover half
Ca)
FIG. 2. Core Wall Model: (a) General Layout; (b) Half Model Analyzed by Finite-Element Method TABLE 1.
Finite-Element Results (Point Load at Top) Shear-lag Shear-lag coefficient coefficient
Model number (1)
2a:2b:H (2)
(3)
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
3:1:5 3:1:10 3:1:20 2:1:5 2:1:10 2:1:20 1:1:5 1:1:10 1:1:20 1:2:5 1:2:10 1:2:20 1:3:5 1:3:10 1:3:20
0.386 0..209 0.. 102 0.301 0.151 0.. 065 0.. 175 0.075 0.035 0..202 0.094 0.. 042 0.222 0.. 107 0.049
TABLE 2. Model number
(1) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
{4)
Stress factor at fixed end (5)
Deflection factor at free end (6)
0.. 313 0.163 0.073 0.290 0.150 0.069 0.267 0.139 0.064 0.427 0.253 0.. 128 0.. 538 0..345 0.. 191
1.299 1..139 1..060 1..255 1.116 1.049 1..221 1..101 1.044 1..432 1..215 1.098 1.657 1..331 1.. 159
1..017 1.006 1.002 1.015 1.005 1.. 002 1.014 1.004 1.002 1.071 1.018 1.004 1.. 162 1.043 1.010
~
Finite-Element Results (Uniformly Distributed Load) Stress Shear-lag Shear-lag coefficient coefficient factor at a. fixed end 2a:2b:H 13 (4) (5) (3) (2)
3:1:5 3:1:10 3:1:20 2:1:5 2:1:10 2:1:20 1:1:5 1:1:10 1:1:20 1:2:5 1:2:10 1:2:20 1:3:5 1:3:10 1:3:20
0.. 533 0..307 0.. 146 0..425 0.220 0..090 0.. 248 0.. 110 0.. 048 0.. 273 0.128 0.. 057 0.. 297 0.144 0.. 066
0.431 0.233 0.104· 0..415 0.. 216 0.100 0.384 0.202 0.096 0..588 0.370 0.. 191 0.707 0..489 0.. 281
1.470 1..217 1..089 1..412 1.179 1..072 1.355 1.156 1.068 1.. 724 1..353 1..155 2.. 100 1.552 1.256
Deflection factor at free end (6)
1..022 1.. 008 1.003 1.020 1..006 1..003 1..018 1..005 1.003 1.. 080 1.021 1..005 1.. 185 1.046 1.012
TABLE 3.
Finite-Element Results (Triangularly Distributed
Load) Model number (1 )
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Deflection factor at free end (6)
Stress Shear-lag She~r-Iag coefficient coefficient factor at 2a:2b:H fixed end a ~ (2) (5) (4) (3) 1.399 3:1:5 0.385 0.478 1.183 3:1:10 0.203 0.267 1.077 3:1:20 0.093 0.126 1.343 2:1:5 0.362 0.378 1.152 2:1:10 0.188 0.193 1.062 2:1:20 0.089 0.075 1.299 1:1:5 0.339 0.218 1.131 1:1:10 0.175 0.094 1.059 1:1:20 0.085 0.042 1.608 1:2:5 0.531 0.247 1.295 1:2:10 0.325 0.114 1.131 1:2:20 0.166 0.050 1.912 1:3:5 0.650 0.267 1.466 1:3:10 0.437 0.130 1.214 1:3:20 0.245 0.059
1.020 1.007 1.003 1.018 1.006 1.002 1.0.16 1.005 1.002 1.077 1.020 1.005 1.177 1.045 1.011
to the web panels in their in-plane directions. Twisting of the channel-shaped structure (half of the .core wall model) is prevented by restraining the free edges of the half flanges (line of· symmetry of the flange panels) from moving horizontally in the in-plane direction. The base of the model is assumed to be perfectly fixed.
Numerical Results The axial-stress distributions in the web and flange panels of a typical model, model 7, which has a flange width/web width ratio of 1.0 and height/half web width and height/half flange width ratios equal to 10.0, are shown in Fig. 3, from y 18014 Eqn. (2)
flange
-----....-x
flange
web
• r Inl te
FIG. 3.
elellant rewl t
Axial Stress Distribution at Base of Model 7
which it can be seen that significant shear lag occurs in both the web and flange panels. The degree of shear lag varies among the models and is dependent on the loading case. To allow detailed study of the shear lag phenomenon, it is proposed to measure shear lag in the web and flange panels in terms of the dimensionless shear lag coefficients, Ctand (3, whose definitions are depicted in Figs. 4 and 5, respectively. These shear-lag coefficients were first proposed by the writer and had been applied to shear lag analysis of framed tube structures in an earlier paper (Kwan 1994). From Fig. 4, which shows the distribution of axial stress across the web panel, it can be seen that when the degree of shear lag is small, the distribution of axial stress across the web panel is approximately linear..-;..-but when the degree of shear lag is large, the axial stress near the centroidal axis of the panel would significantly lag behind that given by a linear distribution leading to a reduction in the gradient of the stress distribution curve at the center in such a way that the larger the shear lag, the greater the reduction. Hence, the reduction in gradient of the stress-distribution curve at the center of the panel may be taken as a measure of the degree of shear lag in the web. The proposed shear-lag coefficient a is defined mathematically as the fractional reduction in· the gradient of the axial stress distribution curve at the center of the web compared to that of a straight line with zero stress at the center of the web and the same maximum stress at the web-flange junction (note that the straight line with zero stress at the center of the web and the same maximum stress at the web-flange junction is not the same as the stress distribution obtained without considering the shear-lag effects because negligence of shear lag would lead to a different value of maximum stress). Similarly, from Fig. 5 which shows the distribution of axial stress across the flange panel, it can be seen that when the degree of shear lag is small, the distribution of axial stress across the flange panel is approximately uniform-but when the degree of shear lag is large, the axial stress near the centroidal axis uf lhe:: panel would lag behind that given by a unifonn distribution leading to a significant reduction in stress at the·. center in such a way that the larger· the shear lag, the greater the reduction. Hence, the reduction in stress at the center of the panel may be taken as a measure of the degree of shear lag in the flange. The proposed shear-lag coefficient ~ is defined mathematically as the fractional reduction in the axial stress at the center of the flange compared to the maximum stress at the weh-flange junction (note that this maximum stress is the maximum stress obtained with shear-lag effects allowed for, not the maximum stress obtained without considering the shear-lag effects). These two shear lag coefficients are dimensionless and are therefore not dependent on the units used. When shear lag is small, the shear-lag coefficients would
e
webflange junction
e
webflange junction
I
I
I
I
I
}-
.
ct (0"
Iv
m
gradient reduced due to shear lag
_I~ } a ~ a
-,lI---L
,
Ca}
FIG. 4. a
(b)
~
a
/
a)
!
,
a
,
}-
~
/
L
'l
(e)
Axial Stress Distributions In Web Panels Illustrating Definition of a: (a) Small Shear Lag; (b) Large Shear Lag; (c) Definition of
JOURNAL OF STRUCTURAL ENGINEERING / SEPTEMBER 1996/1099
.GDiP?i
,
web~
I
flange Junction
b
b
~
I
~ I·
I·
~
stress reduced due to shear lag
J
b
Ca)
FIG. 5. afp
Junction
I
I
D:t::sJ I· J )
\leb-
flange
b
j
~
b
)
b
~
(e)
(b)
Axial Stress Distributions in Flange Panels Illustrating Definition of~: (a) Small Shear Lag; (b) Large Shear Lag; (c) Definition
be close to zero, and when shear lag is large, they would be close to unity. The finite-element results show that the values of Ct and f3 vary along the height of the wall structures and are generally greatest at the bases, Le., the fixed ends, of .the structures. Therefore, the shear lag effects are most critical at the bases. The values of a and 13 at the bases of the models, as determined from the finite-element results, are given in the third and fourth columns of Tables 1-3. From these results, it can be seen that ex is dependent mainly on the heightlhalf web width ratio, while 13 is dependent mainly on the heightlhalf flange width ratio. When the height-ta-width ratios of the panels are small, the values of a and f3 can be as large as or even larger than 0.5 and 0.7, respectively. Comparing the shear-lag coefficients under different load cases, it can also be seen that the importance of shear lag increases in the order of point load case, triangularly distributed load case, and uniformly. distributed load case. Regarding the axial-stress distributions in the web panels, most researchers neglected the shear lag in the webs and assumed a linear distribution of axial stresses across the width of the webs, except Coull and Bose (1975) and Kwan (1994), who considered shear lag in the webs and assumed that the deviation of the axial stresses in the webs from linear distributions can be expressed as third-order-polynomial functions. The present finite-element results, however, reveal that the deviation of the axial stresses in the webs from linear distributions may be more accurately represented by fifth-order polynomials. In Fig. 3, a fifth-order-polynomial curve, whose equation is given by
is plotted on the web panel alongside the finite-element results to demonstrate the close agreement between the axial stress distribution in the web panel and the fifth-order polynomial. Regarding the axial stress distributions in the flanges, the stress distributions across the width of the flanges were approximated as second-order-polynomial curves by Coull and Bose (1975), Coull and Abu EI Magd (1980), and Kwan (1994); as third-order-polynomial curves by Evans and Taherian (1977); and as fourth-order polynomial curves by Moffatt and Dowling (1975). Hence, there is no agreement between the different researchers on the stress distribution in the flange panels. The present finite-element results indicate that the stress distributions in the flange panels lie somewhere between a third-order polynomial and a fowth-order polynomial, but are generally closer to a fourth-order polynomial. A fourth-order polynomial curve, whose equation is given by
CT/(Y) =
el..
[(l - /3) + /3 (~r]
(2)
is plotted on the flange panel alongside the finite-element results in Fig. 3 to demonstrate how close the axial-stress distribution in the flange panel is to a fourth-order polynomial. Shear lag increases the axial stresses at the web-flange junctions and the lateral deflections of the structure. Such effects may be quantified in terms of a stress factor As and a deflection factor Ad as follows: A .r -
axial stress at web-flange junction with shear lag axial stress at web-flange junction without shear lag lateral deflection with shear lag
Ad
= lateral deflection without shear lag
(3)
4 ( )
Since the axial stresses are largest at the fixed ends while the lateral deflections are largest at the free ends, the stress factors at the fixed ends and the deflection factors at· the free ends are more important than those at other locations, and they are tabulated in the last two columns of Tables 1-3. From the tabulated results, it can be. seen that in the worst case of a short and wide wall structure subjected to uniformly distributed lateral loads, the stress factor at the fixed end can be larger than 2.0. The stress factor qecreases as the height of the structure increases, and when the height/half web width and heightJhalf flange width ratios are both greater than 40.0, the increase in bending stress due to shear lag becomes insignificant. On the other hand, the effects of shear lag on lateral deflections are generally much smaller. The increase in lateral deflection due to shear lag would become negligible when the heightlhalf flange width ratio is greater than 10.0.
Comparison with Others' Results The stress factors obtained from the present parametric study are compared to those obtained by Moffatt and Dowling's method (1975) and by Evans and ):aherian's method (1980) in Tables 4 and 5 for the load cases of point load at top and uniformly distributed load, respectively. There are no existing results for triangularly distributed loads and, thus, no similar comparison can be made for this load case. Moffatt and Dowling's (1975) results were obtained by finiteelement analysis, while Evans and Taherian's (1980) results were obtained by the three-bar method. In both Moffatt and Dowling's (1975) results and Evans and Taherian's (1980) results, the effects of any shear lag in the webs were neglected. Comparison of the present results with these" existing results reveals that when the flange width/web width ratio is equal to or greater than 2, there is very close agreement in the stress
1100 I JOURNAL OF STRUCTURAL ENGINEERING I SEPTEMBER 1996
.:~ .... ::,-..~ -'$J~~ .~ ~" ... 1....'/"
:- .•. ~:
.'P
·
TABLE 4. Top)
Comparison with Others' Results (Point Load at Stress Factor at Fixed End
(3)
Present analysis (4)
Moffatt and Dowling (1975) (5)
Evans and Taherian (1980) (6)
1/3 1/3 1/3 1/2 1/2 1/2 1 1 1 2 2 2 3 3 3
1.299 1.139 1.060 1.255 1.116 1.049 1.221 1.101 1.044 1.432 1.215 1.098 1.657 1.331 1.159
1.111 1.064 1.036 1.136 1.078 1.044 1.176 1.099 1.055 1.394 1.207 1.110 1.631 1.321 1.168
1.056 1.028 1.014 1.075 1.038 1.019 1.119 1.059 1.030 1.359 1.179 1.090 1.675 1.338 1.169
Model number (1)
2a:2b:H (2)
bla
1 2 3 4 5 6 7 8 9 10 11 12 13 . 14 15
3:1:5 3:1:10 3:1:20 2:1:5 2:1:10 2:1:20 1:1:5 1:1:10 1:1:20 1:2:5 1:2:10 1:2:20 1:3:5 1:3:10 1:3:20
TABLE 5. Comparison with Others' Results (Uniformly Distributed Load) Stress Factor at Fixed End Model number (1 )
2a:2b:H (2)
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
3:1:5 3:1:10 3:1:20 2:1:5 2:1:10 2:1:20 1:1:5 1:1:10 1:1:20 1:2:5 1:2:10 1:2:20 1:3:5 1:3:10 1:3:20
bla (3)
Present analysis (4)
Moffat and Dowling (1975) (5)
Evans and Taherian (1980) (6)
1/3 1/3 1/3 1/2 1/2 1/2 1 1 1 2 2 2 3 3 3
1.470 1.217 1.089 1.412 1.179 1.072 1.355 1.156 1.068 1.724 1.353 1.155 2.100 1.552 1.256
1.190 1.099 1.058 1.238 1.121 1.071 1.316 1.156 1.090 1.699 1.378 1.182 2.092 1.562 1.290
1.106 1.054 1.028 1.143 1.073 1.037 1.228 1.116 1.059 1.674 1.348 1.177 2.249 1.650 1.331
factors. When the flange width/web width ratio is less than 2, however, the stress factors obtained by the present finite-elementanalysis are significantly higher than those obtained by Moffatt and Dowling's (1975) method or by Evans and Taherian's (1980) method. Such discrepancies may be attributed to· shear lag in the webs. From the differences in stress factors when the flange width/web width ratio is small, it can be seen that shear lag in the webs can contribute to a further increase in maximum bending stress of more than 20% when the webs· are short and wide.
SIMPLIFIED METHOD FOR PRACTICAL DESIGN APPLICATIONS Estimation of Shear Lag Coefficients
The .shear-jag .coefficients, a and 13, are plotted against Hla and Rib, respectively, for. the three loading cases in Figs. 68. From the plotted results, it can be clearly seen that a. is dependent mainly on Hla, while 13 is dependent mainly on HI b.The ratio bla does affect the values of ex and ~, but its effects are gener~lly. mi?or. As et and J3 define the shapes of the axial-stress. dls~butlons, it may be said that the shape of the shear lagged axIal stress distribution in an individual web
.._" .... _- .. ..... :-
....::...~.:_
.. -.
0.& , . . - - - - - - - - - - - - - - - - - - - - - -
•
a
bI.-IIJ
TABLE 6. Fixed End
0.6
Empirical Formulas for Shear Lag Coefficients at Shear lag coefficient
Load
• bfa-I
004
case
ex
(1 )
(2)
Point load at top
•
-
ex=
Prq:loI:cd
1DnnuIa
8
Uniformly distributed load Triangularly distributed load
0.2
Shear lag coefficient J) (3)
1.50 1.00
+ 0.76
1.25 (Hla)
1.59 ex= 1.00 + 0.54 (Hla) 1.56 ex= 1.00 + 0.62 (Hla)
J)
= 1.00 + 0.37
(Hlb)
1.31 J) = 1.00 + 0.24 (Hlb) 1.29 J) = 1.00 + 0.28 (Hlb)
O-+-------+-------+--------li---------f
o
10
20
30
40
Hla
0.1.,.-----------------------. P
•
bla-ll3
•
blplJ2
•
blpl
Point Load at Top
0.6
Model number (1)
a blP2 o bla=a3
0.4
-Prq:loI:cd formula
0.2
0-+-------+-------+----~1--------4
o
20
10
30
40
BIb
FIG. 8.
Values of a and
IJ for Triangularly Distributed Load
or flange panel is quite independent of the dimensions of the other panels. To simplify the estimation of the shear-lag coefficients, it is proposed to neglect the influence of bla on a. and ~, and take the values of ex and ~ as those corresponding to the case of bla = 1.0. Hence, the curves fitting the finite-element results for the case of bla = 1.0 may be taken as design charts for the estimation of ex and ~. Furthermore, to allow quick evaluation using hand-held calculators, formulas for these design charts are derived by empirically matching the finite-element results with different forms of equations. Good matching is found to be achieved with equations of the following forms: CI
a=---C2 + Hla
TABLE 7. Comparison of Stress Factors at Fixed End Obtained by Proposed Formulas with Those by Finite-Element Analysis
Finiteelement analysis (2)
1 2 3 4
1.299 1.139 1.060
5
6 7 8 9 10 11 12 13 14
1.116 1.049 1.221 1.101 1.044 1.432 1.215 1.098 1.657 1.331
15
1.159
1.2S5
Uniformly Distributed Load
Proposed formulas (3)
Finiteelement analysis (4)
Proposed formulas (5)
1.295 1.149 1.075 1.249 1.125 1.063 1.226 1.114 1.057 1.460 1.235 1.119 1.704 1.361 1.183
1.470 1.217 1.089 1.412 1.179 1.072 1.355 1.156 1.068 1.724 1.353 1.155 2.100 1.552 1.256
1.462 1.233 1.117 1.394 1.198 1.099 1.363 1.183 1.092 1.749 1.379 1.191 2.166 1.589 1.296
Triangularly Distributed Load Finiteelement analysis (6)
Proposed formulas (7)
1.399 1.183 1.077 1.343 1.152 1.062 1.299 1.131 1.059 1.608 1.295 1.131 1.912 1.466
1.214
1.390 1.197 1.099 1.332 1.167 1.084 1.306 1.154 1.078 1.630 1.319 1.161 1.971 1.494 1.250
for simplicity, that (j w and CTf can be approximated by (1) and (2). Substitution into (7) and then integration yields 0'",[(4/3)t",a2 (1 - 4na.)
+
4tf ab(1 - 4/5~)] = M
(8)
From this equation, the maximum bending stress is obtained as Ma
(9)
where I w and I, are given by I w = (4/3)tw a 3
(5)
(10)
(11) (6)
where Cl to C4 = unknown coefficients to be determined. The empirical formulas so derived are tabulated in Table 6, and plotted in Figs. 6-8 to demonstrate their close agreement with the finite-element results.
and are actually the moment of inertia of the webs and flanges when there is no shear lag. The effects of shear lag are now clear. Shear lag in the webs causes reduction of the effective moment of inertia of the webs to 1 - O.57a. times the original value, while shear lag in the flanges causes reduction of the effective moment of inertia of the flanges by a factor of 1 0.80(3. From (9), the stress factor is evaluated as
Estimation of Stress Factors The bending stresses are governed by the following moment-equilibrium equation:
f:a
2tw CTw X dx
+
f:'
2r,CT,a dy = M
(7)
in which t w and ~ = thickness of web and flange panels; and M = bending moment acting on section. It may be assumed, 1102/ JOURNAL OF STRUCTURAL ENGINEERING / SEPTEMBER 1996
As
[w + I, =-------~---[w(1 - 0.57a) + [,(I - 0.80~)
(12)
The stress factors so evaluated by using the empirical formulas given in Table 6 and the prior equation are compared with those obtained by finite-element analysis in Table 7. It is seen that within the range of parameters studied, the stress factors evaluated by the proposed fonnulas differ by, at most, a few percent from those by finite-element analysis. Hence, the pro-
-
;- ;;
~_
.
.
20
posed formulas are considered sufficiently accurate for practical applications.
fa
tf
10
ANALYSES
III
It
5
n
III
180 JcH/1i
n
-load
(b) II
3.62 HPa
i
I
1
90 leN/,.
3.62 HPa
\leb {lanalt
Ca)
(e)
FIG. 9. Example: (a) Core Wall SUbjected to Wind Load; (b) Cross Section; (e) Axial Stress at Base As Obtained by Finite-Element Analysis
high accuracy, the variation of thickness with height may be neglected in the evaluation of the degree of shear lag at the base. The value of Poisson ratio used in the analysis would also affect the numerical results for the shear-lag effects. There is, however, the problem of what Poisson ratio should be used in the analysis. Concrete has a Poisson ratio under static load of 0.15-0.20, but an average Poisson ratio under dynamic load of about 0.24 (Neville 1981). The presence of reinforcement affects the Poisson ratio, too. Although reinforced concrete is not homogeneous, it is often treated as a homogeneous material with equivalent properties in order to simplify the structural analysis. This can be done by smearing the reinforcement across the concrete section and taking the effective Young's modulus E of the reinforced concrete as E = (1 - p)Ec
+
pE.r
(13)
where Ec and E.r = Young's moduli of concrete and steel, respectively; and p = reinforcement ratio. However, since reinforced concrete walls can be deformed in shear without straining the horizontal and vertical reinforcement, the effective shear modulus G should be unaffected and remain the same as that of plain' concrete. For compatibility, the effective Pois.. son ratio of the reinforced concrete needs to be taken as that evaluated' by 2(1+ v) = EIG
Limits of Applications In the present study, homogeneous, isotropic walls of constant thickness and a Poisson ratio of 0.25 have been assumed. Because of the assumption made regarding homogeneity, the proposed method should not be applied to precast concretepanel structures whose joints, as planes of weakness, could render the structural behavior quite different from that of homogeneous walls. Theoretically, anisotropy in reinforced concrete walls due to differences in horizontal and vertical reinforcement could affect the stress distributions. But, as the lateral stresses in the walis are normally quite small compared to the longitudinal stresses, it is unlikely that the Young's modulus in the lateral direction would significantly affect the overall. structural behavior. Thus, the effects of anisotropy may be neglected and the present results should be applicable to reinforced concrete walls albeit they may not be entirely isotropic. Strictly speaking, variation of wall thickness with height may affect the shear-lag .phenomenon. Nevertheless, since shear lag is basically a local-stress-concentration .problem at the lower part of the wall structure, it· is anticipated' that the wall thickness at. the upper part of the structure would not significantly affect the shear lag at the base where it is most critical. Therefore, for preliminary designs that do not require
:: thickness ft or vall. :: - 0.3 III
ft Of
If
50
In actual practice, the applied lateral loads are never as simple as the three loading cases considered. Nevertheless, for the purpose of estimating the shear-lag coefficients, the actual loading case may be approximated, by exercising engineering judgment, as one of the three loading cases studied. An example of a core wall subjected to wind loads is given in Fig. 9. In this case, the applied loads consist of several uniformly distributed loads of increasingly larger intensity at greater height. Although the loads are not distributed as in any of the three loading cases studied, the load distribution is treated, for the purpose of estimating a. and ~, as a triangular distribution. Using the proposed formulas for the triangularly distributed load case, the values of a. and (3 at the base of the core wall are evaluated as 0.380 and 0.339, respectively. The bending moment acting at the base of the core wall is then calculated from the actual load distribution as 264.4 MNm. Finally, the maximum bending stress at the base is obtained by using (9) as 3.52 MPa. For comparison, finite-element analysis of the core wall using the actual load distribution is carried out and the bending stress results are shown in Fig. 9(c). The maximum bending stress obtained by finite-element analysis is 3.62 MPa, which is very close to that evaluated previously by hand calculation. Moffatt and Dowling (1975) suggested that for bridges, if the applied loads are type HA or HB vehicle loads, as defined in British Standard BS5400 ("Steel" 1978), for the .purpose of estimating the reduced effective widths of the flanges due to shear lag, a uniformly distributed load case may be assumed. For buildings, a' similar assumption may be made. If the applied loads are wind loads whose distribution is similar to that in the previous example, then a triangularly distributed load case may be assumed in the evaluation of the shear-lag coefficients. When the load distribution lies between unifonnand triangular distributions, then the mean value of the shearlag coefficients for uniformly and triangularly distributed load cases may be used.
III
U- l o a d
5IU
270 kN/1ll
Load Case
20
I
(14)
from which a higher Poisson ratio than that of plain .concrete would be obtained. For typical walls cast of concrete having a Poisson ratio of 0.15-0.20 and with 0.3-1.2% reinforcement provided, the effective Poisson ratio ranges from 0.18 to 0.30. Under dynamic load, the Poisson ratio would be slightly higher. In actual engineering practice, however, it is not really necessary to be so precise in the evaluation of the Poisson ratio. For most applications, it should be sufficiently accurate to just use an average value of 0.25 regardless of the amount of reinforcement provided, whether the load is static or dynamic.
CONCLUSIONS A parametric study of the shear-lag phenomenon in shearl core wall structures has been carried out by analyzing a number of core wall models with.the finite.;.element method. In the study, shear lag in allthe web and flange panels is taken into account, and the load cases considered include point. load at top, unifonnly distributed load, and triangularly distributed loads. The numerical results showed that (1) the degree of shear lag in a cantilevered wall structure varies along the height and is generally greatest at the fixed end; (2) the. axial stress disJOURNAL OF STRUCTURAL ENGINEERING I SEPTEMBER 1996/1103
tributions across the widths of the web and flange panels can be described approximately by polynomials of fifth and fourth order, respectively; (3) the importance of shear lag increases in the order of point load case, triangularly distributed load case, and uniformly distributed load case; and (4) the effects of shear lag in the web panels can be quite significant when the web panels are relatively short and wide, and hence, it should be prudent to also take into account the effects of any shear lag in the webs. Detailed analysis of the shear-lag phenomenon revealed that the degree of shear lag in an individual web or flange panel, measured in tenns of the dimensionless shear-lag coefficients a or J3, is dependent mainly on the height/width ratio of the panel. Plotting the shear-lag c;oefficients against the heightJ width ratios of the panels and matching the numerical results with empirical equations of different forms, design charts and empirical formulas for estimating the shear-lag coefficients are produced. A simple equation for evaluating the increase in maximum-bending stress due to shear lag in both the web and flange panels is also derived. Comparison with the finite-element. results confirmed that the proposed formulas are sufficiently accurate for practical applications. A numerical example has also been presented to demonstrate the ease of application of the proposed formulas.
APPENDIX I.
REFERENCES
Abdel-Sayed, G. (1969). ""Effective width of steel deck-plate in bridges." J. Struct. Div., ASCE, 95(7), 1459-1474. Connor, J. J., and Pouangare, C. C. (1991). ""Simple model for design of framed-tube structures." J. Struct. Engrg., ASCE, 117(12), 3623-
3644. Coull, A., and Abu EI Magd, S. A. (1980). Analysis of wide-flanged shearwall structures." Reinforced concrete structures subjected to wind and earthquake forces, ACI Spec. Publ. 63, Paper No. SP63-23, Concrete Institute, Detroit, Mich., 575 -607. Coull, A., and Bose, B. (1975). "'Simplified analysis of frame-tube structures." J. Struct. Div., ASCE, 101(11), 2223-2240. DeFries-Skene, A., and Scordelis, A. C. (1964). "'Direct stiffness solution for folded plates." J. Struct. Div., ASCE, Vol. 90(4), 15-48. Evans, H. R., and Taherian, A. R. (1977). '''The prediction of shear lag effect in box girder." froc., Ins tn. Civ. Engrs., Part 2, 63, Thomas Telford Services Ltd., London, 69-92. Evans, H. R., and Taherian, A. R. (1980). ""A design aid for shear lag calculations." Proc., Instn. Civ. Engrs., Part 2, 69, Thomas Telford Services Ltd., London, 403 -424. Kristek, V. (1979). ""Folded plate approach to analysis of shear wall systems and frame structures." Proc., Instn. Civ. Engrs., Part 2, 67, Thomas Telford Services Ltd., London, 1065 -1075. Kristek, V. (1983). "Chapter 6: Shear lag in box girders." Plated U
1104 / JOURNAL OF STRUCTURAL ENGINEERING / SEPTEMBER 1996
structures-stability and strength, R. Narayanan, ed., Applied Science Publishers, London and New York, N.Y., 165-194. Kristek. V. and Skaloud, M. (1991). Advanced analysis and design of plated structures-Developments in Civil Engineering 32. Elsevier Science Publishing Co. Inc., New York., N.Y. Kwan, A. K. H. (1994). "Simple method for approximate analysis of framed tube structures. u J. Struct. Engrg., ASCE, 120(4), 1221-1239. Moffa~ K. R., and Dowling. P. J. (1975). "Shear lag in steel box girder bridges." Struet. Engrg., 53(10), 439-448. Neville, A. M. (1981). "Chapter 6: Elasticity. shrinkage and creep." Properties of concrete. Longman Scientific and Tech. Ser., Longman Group U.K. Ltd.• Essex, England, 359-432. Song, Q., and Scordelis, A. C. (1990a). hFonnulas for shear-lag effect of T-, 1- and box beams." J. Struct. Engrg., ASCE, 116(5), 13061318. Song. Q., and Scordelis A. C. (1990b). "'Shear lag analysis of T-, 1- and box beams:' J. Struct. Engrg., 116(5), 1290-1305. "Steel, Concrete and Composite Bridges." (1978). BS5400, British Standards Instn., London. Taherian, A. R., and Evans, H. R. (1977). uThe bar simulation method for the calculation of shear lag in multi-cell and continuous box girders." Proc., lnsm. Civ. Engrs., Part 2, 63 Thomas Telford Services Ltd., London, 881-897. 9
9
9
APPENDIX II.
NOTATION
The following symbols are used in this paper: a = half width of web panel; b = half width of flange panel; C, = coefficients; E = effective Young's modulus; Ec = Young's modulus of plain concrete; E s = Young's .modulus of steel reinforcement; G = shear modulus of concrete; H = height of wall; IJ = moment of inertia of flange; I w = moment of inertia of web; M = bending moment; ~ = thickness of flange; tw = thickness of web; a = shear lag coefficient of web panel; ~ = shear lag coefficient of flange panel; Ad = deflection factor as defined by (4); AI = stress factor as defined by (3); v = Poisson's ratio; p = reinforcement ratio; a', = axial stress in flange panel; am = axial stress at web-flange junction (maximum stress in section); and a'w = axial stress in web panel.