You are on page 1of 32
CHAPTER 9 Unsymmetric Bending, Shear Center, and Thin-Walled Beams Bending stresses are analyzed in solid and thin-walled straight beams whose cross sections have no axis of symmetry. Load placement for bending without twisting is considered. Simplified analysis methods are presented for beams with very thin webs. AREA MOMENTS AND PRODUCTS OF INERTIA In the elementary flexure formula @ = Mc/I, the symbol / represents the mo- ment of inertia of the cross-sectional area of the beam about the neutral axis. In subsequent discussions we need also the product of inertia. Moments and products of inertia of a plane area are briefly reviewed in this section. More detailed discussions appear in texts on statics. Definitions. The symbols /,, /;, J, and /,, have the respective names moment of inertia about the s axis, moment of inertia about the ¢ axis, polar moment of inertia about point O, and product of inertia with respect to the s and t axes. With reference to an arbitrary plane area and arbitrarily placed axes s and 1, Fig. 9.1.1a, we have by Uefinition 345 346 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS FIGURE 9.4.1. (a) Arbitrary plane area A, with differential clement dA. (b) Area symmetric about the ¢ axis. (c) Centroidal axes yz and distances d, and d. used in parallel axis theorems. (d) Orthogonal axis systems st and én, having common origin but different orientation. =f Pda, I, -{ sda, eee area (9.1.1) Jo = for dA, 1a = [. st dA Since r? = s? + #°, we see that J, = 1, + J,. Although /,, /,, and J, are always positive, the product of inertia /,, can be positive, negative, or zero. If s or sis a symmetry axis of area A, then /,, = 0. The argument is shown in Fig. 9.1.1b, where ¢ is an axis of symmetry. Each positive contribution + st dA is matched by an equal negative contribution ~st dA, Summing over A, we obtain 1, = 0. Transfer of Axes. Transfer of axis theorems relate quantities in Eqs. 9.1.1 to corresponding quantities referred to parallel axes whose origin is at the cen- troid of the area. In the notation of Fig. 9.1.1c, the transfer theorems are t, =f, + Ad? J, = Ig + Ar* 1,=1,+ Ad? 1, =1,, + Add, a (9.1.2) where f,, ,, Jg, and /,, are the moments and product of inertia of area A about centroidal axes yz. Distances d, and d, are signed quantities. For example, if point G were below the s axis, d, would be negative. All four theorems are proved in similar fashion. For example, the proof of the last one is A= | stdA = I (y + dle + d.) dA area area ees [ooveaa +a [eda +a [yaa + ad, [an (9.1.3) + 0+ 0+ Add, = I, + Add, The integrals of z dA and y dA vanish because they represent first moments of the area about centroidal axes yz. 9.1 AREA MOMENTS AND PRODUCTS OF INERTIA 347 Principal Axes. For a given area A and a given origin O, there are two axes about which the moments of inertia are, respectively, greater and less than about any other axes through origin O. These are the principal axes. Inertias Im, and Imin Teferred to them are the principal moments of inertia. These quantities are found as follows. e In the notation of Fig. 9.1.1d, te= |. eda = | (t.cos @ — s sin 6) dA ” i= ga = | (1 sin 6 + s cos 0)? dA (9.1.4) Len = gran = | (t cos @ — s sin 6)(t sin 8 + s cos 6) dA ea area After integration and use of trigonometric identities for sin?@ and cos?6, there results I, = 40, + 1) + Md, — 4) cos 26-1, sin 20 1, = Ml, + 1) — Mi, - 1) cos 26 + 1, sin2@ ** (9.1.5) Teq = Hd, ~ 1) sin 2 + 1,, cos 20 < The value of @ that extremizes I, and I,, is found from di,/d@ = 0 or from dl,/d@ = 0. This angle, the principal angle 6,,, is defined by tan 26,, = (9.1.6) Im ds Angle 8, has two values, 7/2 apart, one for In,, and the other for mim. BY using Bq. 9.1.6 in Eqs. 9.1.5, we find the principal moments of inertia: ff-1\. Frnaxin = h > he i 4) +2, (9.1.7) This equation easily yields Imag + Lin = Ay + Lp Uf Finax = Iminy all axes are principal, as is clearly true for centroidal axes of a circular cross section. Equations 9.1.5 have the same form as plane stress transformation equations. I,, and /,, are analogous to normal and shear stresses. A Mohr’s circle for /’s can be drawn, like the circle for stresses in Fig. 1.4.2. Details of the procedure are explained in textbooks on statics. Since the product of inertia is analogous to shear stress, and shear stress vanishes on principal planes, Ie, = 0 (when & and 7 are principal axes) (9.1.8) That this is so can be seen by noting that Uz, = d/,/d0, so Iz, must vanish when 6 = 6,,. If Jz, = 0 for orthogonal axes én, these axes are principal. Accordingly, if £or 7 is an axis of symmetry, then £ and 7 are principal axes. 348 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS Example 9.1.1, For the cross section shown in Fig. 9.1.2, find the mo- ments and product of inertia about horizontal and vertical centroidal axes. Also locate the principal centroidal axes and find the principal centroidal moments of inertia. By the usual calculations explained in textbooks on statics, the centroid is found to be point O. The cross section is arbitrarily divided into parts 1 and 2 for the following calculations. Applying Eqs. 9.1.2 gives us _ [20(100)3 60(20)3 a 12 2 at 20005] + | + 120257 (9.1.9) where the two bracketed expressions come from parts 1 and 2, respectively. A similar calculation yields /,, and /,, is 1,. = [0 + 2000(- 15)(—15)] + [0 + 1200(2525)] (9.1.10) Collecting results, we have 1, = 2.907(10)° mm* J, = ‘1.627(10)° mm* = 1.200(10)° mm* (9.1.11) From Eq. 9.1.6, 201.200) tan 20m = T 607 — 2.907 8, = —31.0° (9.1.12) which is the clockwise angle shown in Fig. 9.1.2. The other possibility, 6,, + 90° = 59°, is the angle between the y and n axes. From Eq. 9.1.7, Igy = 3-627(10)° mm? —Iyin, = 0.907(10)° mm* (9.1.13) In this example it is clear by inspection of Fig. 9.1.2 that I, = Imax rather than J, = Imax. In less obvious cases the two angles 4 can be substituted into Eqs. 9.1.5, and the larger of /,, and /,, identified as J,,,,. Alternatively, the Mohr’s circle construction suggested above Eq. 9.1.8 immediately identifies the larger of J, and /,. Algebraic signs are important in the calculation of /,, and 0,,. Notice that /,, can be positive or negative for a given shape, depending on its orientation (| 9.1.3). We see that /,, of the triangle is negative when most of its area lies in the second and fourth quadrants. 20mm 80 mm a) “ales min FIGURE 9.1.2. Angle cross section. Axes - bso mm: yz and mn are centroidal, and axes mn are also principal. 9.2 UNSYMMETRIC BENDING OF STRAIGHT BEAMS 349 FIGURE 9.1.3. Product of inertia of a right tri- angle. Axes y and z are centroidal. 9.2 UNSYMMETRIC BENDING OF STRAIGHT BEAMS Elementary theory of bending of a straight beam is reviewed in Sections 1.1 and 1.2, Elementary theory presumes that the beam has at least one plane of symmetry and that transverse loads on the beam act in a plane of symmetry. Such a case is shown in Fig. 9.2.1. Bending stress is given by the familiar flexure formula, as shown in Fig. 9.2.1c. The neutral axis, defined by a, = 0, is parallel to the moment vector and perpendicular to the plane of loads. Things are not as simple if the cross section has no plane of symmetry, as in Fig. 9.2.2a. Bending of such a beam is called unsymmetric bending. The fol- lowing discussion develops general expressions for orientation of the neutral 2 Be at a\ ® 8 he We FIGURE 9.2.1. Elastic flexural stress in a beam where the xz plane is a plane of symmetry.,The loads produce mo- ment M,, which acts on material behind the vz plane. Vector M, is parallel to the y axis, which is a principal axis of the cross-sectional area, Point O is the centroid of the cross sec- tion, © 350 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS: Neutral axis EEE SED ST] pe 7 LF M FIGURE 9.2.2. Beam whose cross section has no axis of symmetry. Loads produce moment M y on cross section ABC. The y and z components of M are M, and M.. In general, A # B. Point O is, the centroid of the cross section: otherwise, axes y and z are arbitrary orthogonal coordinates. oy axis and flexural stress o, in a straight beam whose cross section has no axis of symmetry. Deflections of the beam are considered in Section 9.3. Expressions to be developed are not restricted to the triangular cross section depicted in Fig. 9.2.2. In Fig. 9.2.2, moment ™ is applied to the x = 0 face of the beam segment that lies behind the yz plane by the beam segment that lies in front of the yz plane. The y and z components of M are M, and M.. In general, for any shape of cross section, M, and M, are moment components applied to a cross section by the portion of the beam that is cut away to expose the cross section. Assumptions made in the following analysis are familiar ones. The material is homogeneous, isotropic, and linearly elastic. The beam is straight and of constant cross section. Flexural stress o, is unaffected by shear stresses that may be present because of transverse load. (Sections 9.6 through 9.9 discuss where load must be placed to avoid twisting.) We begin by considering the deformation produced by bending. Whether or not there is a plane of symmetry, experiment shows that plane sectlons remain plane (except where “‘shear lag’’ is significant; see Section 9.14). Cross sections rotate about a neutral axis whose location and orientation are as yet unknown. This implies that axial strain ¢, is a linear function of y and z. As is usual (and 9.2 UNSYMMETRIC BENDING OF STRAIGHT BEAMS 351 appropriate) in beam theory, we assume that normal stresses o, and , are either zero or negligible in comparison with axial stress o,. Accordingly, 7, = Ee, and a, is also a linear function of y and z. O= at by +cz (9.2.1) where a, b, and c are constants. To find them, we write equations of equilibrium. o={ odd M, -{, z0,dA M, = -f yo, dA (9.2.2) In these expressions, 0, dA is a differential force dF produced by a, acting on differential area dA (Fig. 9.2.2b). The first integral says that since there is no axial load on the beam, differential forces dF sum to zero. The second integral says that differential moments z dF about the y axis sum to M,. The third integral imilar to the second, and requires a negative sign because, when y and o, are both positive, the differential moment yo, dA is opposite in direction to M, in Fig. 9.2.2b. Next, we substitute Eq. 9.2.1 into Eqs. 9.2.2. Because axes y and z are centroidal, fy dA = fz dA = 0. Also, by definition, 1, -{ 2dA |, =| area area which are the moments and product of inertia of the cross-sectional area with respect to y and z axes. Solving for a, b, and c, we find a = 0, which means that the neutral axis passes through the centroid y = z = 0. Substituting the solutions for b and c into Eq. 9.2.1, we find the flexural stress. I. yzdA (9.2.3) (-Myl,, — M,I)y + My, + Maly)? Il. — 2 (9.2.4) (See Eq. 9.3.2 for an alternative form of the expression for o,.) On the neutral axis, 0, = 0. To locate the neutral axis, we set o, = 0 in Eq. 9.2.4 and find, for angle A in Fig. 9.2.2c, tan A = (9.2.5) Coordinate system yz is of arbitrary inclination with respect to sides of the cross section but must be centroidal. Remarks. If the beam carries an axial force N whose line of action is through centroids of cross sections, it produces uniform axial stress 0, = N/A, which must be superposed on the flexural stress of Eq. 9.2.4. The presence of N translates the neutral axis away from the centroid but does not change its orientation. Remember that Eqs. 9.2.4 and 9.2.5 require that the beam be homogeneous. and linearly elas' 352 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS If y and z are principal centroidal axes, then /,, = 0, and Eq. 9.2.5 reduces to (when J, = 0) (9.2.6) Thus, we find o,, by superposition, using the elementary flexure formula about each principal centroidal axis of the cross section. Clearly, Eq. 9.2.6 is simpler than Eq. 9.2.4, and is usually preferred when principal centroidal axes can be located by inspection, as for a cross section with an axis of symmetry. The resultant moment vector M and the neutral axis are parallel if and only if M is parallel to a principal axis of the cross section. Example 9.2.1. The simply supported beam in Fig. 9.2.3a has an angle cross section whose dimensions are given in Fig. 9.1.2. (Properties of this cross section are evaluated in Example 9.1.1.) Locate the neutral axis. De- termine where the largest tensile and compressive flexural stresses appear, and calculate them in terms of load Q. The load is parallel to the z axis, so M, = 0 at all cross sections. From Eqs. 9.1.11 and 9.2.5, 1 1.200 = ee —— = 36.4° 9.2.7, A = arctan 1 arctan 1.627 6. ( ) A sketch to scale, Fig. 9.2.3b, immediately shows that A and B are the points most distant from the neutral axis. Stress varies linearly with distance from Q(N) Q 907 (10) mm4 627 (10) mm4 300 110) mnt FIGURE 9.2.3. (a) Simply supported beam with an angle cross section whose dimensions are given by Fig. 9.1.2. (b) Neutral axis location and re- gions of tensile and compres- © sive flexural stress. 9.2 UNSYMMETRIC SENDING OF STRAIGHT BEAMS 353 the neutral axis. Therefore, points A and B beneath load Q carry the extreme stresses. Point A has coordinates y = —25 mm and z = 35 mm. With M, = —0.69(1000) = —600Q N-mm and M, = 0, Eq. 9.2.4 yields [-1.200(-25) + 1.627(35)]10° Sa = ——3 98 8(10)'@ «( 8008) = ~G01592 MPa 0.2.8) where 3.288(10)'? mm* = JJ, ~ /2,. Similarly, at point B, where y = —5 mm and z = —65 mm, [=1.200¢-5) + 1.627(-65)]10° Ga = 93 (= 6000) = 0.01820 MPa (9.2.9) As should be anticipated on physical grounds, o,, is compressive and oy is tensile. What is unexpected, from the limited viewpoint of symmetric bending theory, is the zone of tensile a, above the y axis. But if a, were entirely compressive above the y axis, moments about an axis that contains load Q would not sum to zero as required for static equilibrium. First Alternative Solution. Instead of using Eqs. 9.2.4 and 9.2.5, we can use Eqs. 9.2.1 through 9.2.3, thus effectively repeating the derivation of Eqs. 9.2.4 and 9.2.5 for the case at hand. There is no axial force, so that a = O in Eq. 9.2.1. Equations 9.2.2 and 9.2.3 yield M, = —6000 = b/,, + cl, (9.2.10) M, = 0 =I, + cl, Inserting numerical values of /,, /., and J,,, we find b = 219.0(10)-°Q = —296.8(10)-°Q (9.2.11) From Eq. 9.2.1, with @, = 0, the neutral axis has the orientation 0 pao ey ee RC From Eq. 9.2.1, with y = —25 mm and 2 = 35 mm, oa = [219.0(-25) — 296.8(35)]10-°9 = —0.01599 MPa (9.2.13) Again from Eq. 9.2.1, with y = —5 mm andz = —65 mm, oy = [219.0(-5) — 296.8(—65)]10-°Q = 0.01829 MPa (9.2.14) Results are of course the same as before. This solution method is no more complicated than the preceding method and has the merit of keeping us more in touch with the physics of the problem. tan dA = Second Alternative Solution. After finding A from Eq. 9.2.5, o can be found from Eq. 9.2.6. Here /, and /, are principal moments of inertia of 354 500 mm BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS: the cross section (already calculated in Eqs. 9.1.13), and M,, M,, y and z become moments and coordinates with respect to principal axes, For exam- ple, with @,, = —31.0°, M, = (— 6009) cos 6, = —514.3Q —(—6009) sin 6,, = —309.09 ya = (—25) cos 8,, + (35) sin 4, — 39.46 mm (9.2.15) z, = (35) cos 8, — (—25) sin 6, = 17.12 mm —309.00( — 39.46) _ —514.3Q(17.12) aan o.s079% ~~ 3.62700)" ~~~ 01992 MPa This solution seems least preferable of the three because of the need to locate principal axes, compute Jina, aNd Z;piny and find new coordinates of the point in question, with the attendant chance for numerical and sign error in every step. (The method may still be a good one for a standard rolled section because @,,. Inax» and /,,;, are readily found from tabulated information.) Example 9.2.2. A maximum bending moment M is developed in a simply supported beam. The plane of loads is inclined at an angle B with respect to the vertical (Fig. 9.2.4a). For 8 = 2°. locate the neutral axis and find the flexural stress at corner C in terms of M. Compute the factor by which this stress is increased by changing f from 0° to 2° for the same M. Here Eq. 9.2.6 is appropriate because principal axes and principal /’s are easily found. From 7 = bh?/12, 2, = 416.7(10)° mm* I, = 2.667(10)° mm (9.2.16) and I, = 0. For B = 2°, from Eq. 9.2.5, the angle between the y axis and Neutral axis, \44 a A B Ty A aA BA 79.6" Be Van We Wy Ba ’ OZ y Yu BY Oy 44) Y yy 4 Ac vAbe Plane of loads “| [mn fa) (b) FIGURE 9.2.4. (a) Deep, narrow rectangular cross section. Axes y and z are prin- cipal centroidal axes. (b) Location of the neutral axis for B 9.3 BEAM DEFLECTIONS IN UNSYMMETRIC BENDING 355 the neutral axis is (=M sin 2°), : (=M cos 2°91 At corner C in Fig. 9.2.4, Eq. 9.2.6 yields, for B = 2°, c= — eM sin 22120) M cos 2°)(~250) aa 2.667(10)® 416.7(10)° (9.2.18) O,c = (0.2617M + 0.5996M)10~° = 0.8614(10)- °M For B = 0, we find o,¢ = Mz/I, = 0.6000(10)~°M. The factor by which Oc is increased is 0.8614/0.6000 = 1.44. The small change from B = 0° to B = 2° has appreciably increased maximum stress and has rotated the neutral axis through a large angle from the horizontal. If the beam is used as a floor joist, the floor provides lateral restraint along the upper edge, effectively making 8 = 0. Beams for which 1, >> I, should be provided with lateral support of this or a similar kind. = 79.6° (9.2.17) BEAM DEFLECTIONS IN UNSYMMETRIC BENDING Again we consider straight beams. In elementary theory, a principal axis of the beam cross section lies in the plane of loads, and the beam deflects in that plane. We use elementary formulas, such as PL?/3EI for the deflection of a straight tip-loaded cantilever beam, where / is the areca moment of inertia about the neutral (and principal) centroidal axis of the cross section. If the plane of loads is not parallel to a principal axis of the cross section, or if there is more than one plane of loads (owing perhaps to both horizontal and vertical forces on the beam), we can proceed as follows. Resolve each load into two components, one parallel to each of the two principal centroidal axes of the cross section, to produce two sets of loads, one in each principal direction. For each set calculate deflections by elementary theory, as described in the preceding paragraph. At each point on the beam axis, the two deflection components are mutuaily perpendicular, and can be added vectorially The foregoing calculations are simple if principal axes are easily found, as for sections with a plane of symmetry. For unsymmetric cross sections the following method is easier. It requires that the neutral axis maintain a constant inclination A along the beam, as happens when there is a single plane of loads and the cross section is constant along the beam. If two or more loads have different inclinations, so that A is not constant, calculations must be done for each load, and the results added vectorially. In Fig. 9.3.1a, A is Weflection of the centroid O (and of any other point on the neutral axis). Deflection A is perpendicular to the neutral axis and has two components, v and w. The y direction component v is negative when A is positive. An expression for A is now derived. 9.5 SHEAR STRESS AND SHEAR FLOW IN STRAIGHT BEAMS 359 FIGURE 9.4,1. End view of a cantilever beam of length L with constant cross section and a constant rate of pretwist @. Axes x and y are principal centroidal axes of the root sec- tion, Axes & and 7 are principal centroidal axes of an arbitrary cross section between x = Oandx = L. Complementary strain energy of bending is u* f Mi » it d (9.4.2 ~ Jo \zer, * 2e1,) nee. Deflections at the tip in the y and z directions, respectively, are au* au* = .=— 4.3 a= 5 F, oF (9.4.3) After tedious manipulation, Eqs. 9.4.1, 9.4.2, and 9.4.3 yield (9.1] a FL (| Gat hie PL * EL, \6 7 El, FLL (1 FL A= [54 eeteteeett a (4) where, with & = aL, c, = 2h sin 2h 20 — 1 + cos 2h ie a ae For the limiting case a = 0, we find C, = ¢ and C, = 0, so that A, = F,L3/3El, and A, = F.L3/3El,, as expected. Reference {9.1 also considers deflection due to uniformly distributed load. If a were not constant, or the beam were tapered or carried a complicated load distribution, one could evaluate U* numerically. For this the beam would be divided into length increments Ax, with all quantities taken as constant over a typical Ax, and Eg. 9.4.2 regarded as a summation over all increments. However, computer methods may be more appropriate for such problems [8.12] Also, the restriction as to shape of cross section can be removed [8.13]. (9.4.5) SHEAR STRESS AND SHEAR FLOW IN STRAIGHT BEAMS On occasion it is necessary to calculate shear stress in a beam because its material is weak in shear, as in a-wooden beam. Also, the distribution of shear stress over the cross section must be understood if we are to locate the shear center, “BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS: as is done in subsequent sections. Formulas developed in this section are limited to beams that are thin-walled, of open cross section, and without taper. For such a beam the only shear stress of any consequence acts along the beam and tangent to the medial line of the cross section. This shear stress is labeled 7,, if s is a coordinate along the medial line. The restriction to nontapered beams is needed because taper reduces transverse shear stress (e.g., in the tip-loaded cantilever wedge of Problem 2.35, transverse shear stress is zero on the neutral axis of bending). We also require that the beam be of one material, subsequent derivations invoke the flexure formula, which presumes homogeneity. However, a beam of two or more materials can be analyzed by using modifications de- scribed in the paragraph that follows Eq. 9.5.6. ‘‘Shear lag” may influence stresses if the beam has wide, thin flanges; see Section 9.14. We begin with a review of concepts and formulas from elementary theory for symmetric open cross sections. Closed sections are considered in Section 9.9. Symmetric Cross Sections. The cantilever beam in Fig. 9.5.1 carries a load Q parallel to the z axis. At each cross section there is a shear force V = Q in the z direction and a bending moment M about the y axis. Shear stress 7 is produced by V. We ask first for the direction of 7 at an arbitrary point in the cross section. Consider face ABC of the differential element in Fig. 9.5.1. Axial force N is produced by axial flexural stress of magnitude 7, = Mz/I. On the opposite face the axial force is N + dN because the bending moment is larger there, having magnitude M + dM. Outer edge BD is free of axial force. The inner face of “the differential element carries axial force dF, which is produced by shear stress +. Force dF must act in the direction shown if axial forces on the differential element are to sum to zero. Shear stress on face ABC must be directed toward +E Section FE dF =gde =r dx FIGURE 9.5.1. Diagrams used to determine the direction of shear stress and derive a for- mula for its magnitude. Axes yz are centroidal principal axes of the cross section. Section BE is viewed looking toward the free end. 9.6 GHEAR STRESS AND SHEAR FLOW IN STRAIGHT BEAMS 361 B (shear arrows on adjacent faces must both point either toward or away from a corner, as shown by elementary theory of stress). These arguments, applied with cuts at various locations on the cross section, serve to establish the direction of 7 throughout section EE. By expressing the foregoing arguments quantitatively, we derive an expres- sion for shear stress r at line AC, which spans the thickness at an arbitrary location. Force N is, with area increment dy dz = dA, v= |. ea ABC I where A, is the area of face ABC and Z is distance between the centroid of A, and the y axis. Thus A,Z is the first moment of area ABC about the neutral axis of bending. In general, A,z pertains to all of the cross-sectional area that lies to one side of the thickness line AC along which 7 is sought. Similarly, forces N + dN and dN are (9.5.1) M + dM N+ dN = AAR and = dN = (9.5.2) For equilibrium of axial forces, N+dF-(N+dN)=0 so dF =dN (9.5.3) Substituting Eqs. 9.5.2 and dF = vt dx, we find dM A,z a 5.4) feereeeoeee (9.5.4) From elementary theory, dM/dx has magnitude V, where V is the transverse shear force on the cross section in question. Defining q as the shear flow g = tt, whose dimensions are force per unit length, we may state the final result in two ways: VAZ IZ Tree cae an (9.5.5) If ¢ is constant, 7 and q vary in the same way over a cross section. Equations 9.5.5 are known ‘as Jourawski’s formulas, after the Russian engineer who in 1844 was the first to give practical formulas for transverse shear stress in beams. Equations 9.5.5 apply to a thin-walled homogeneous beam without taper where y and 2 ate principal axes of the cross section. Shear stress t acts axially along the beam as well as tangent to the medial line of the cross section and is assumed to be constant across the thickness ¢. The other two shear stresses (such as 7,, and 7,, in the stem of the T section) must vanish on surfaces of the beam, and therefore must be small at interior points. We assume that they are negligibly small. These assumptions are nearly exact for a thin-walled cross section. (If applied to a rectangular cross section of width b and depth h, Eqs. 9.5.5 produce the familiar result 7 = 3V/2bh. According to Section 124 of [2.1], this result is appreciably in error if b > h, being low by a factor of 5.2 if b = 15h.) 362 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS Note that the shear stress calculated in Eqs. 9.5.5 has nothing to do with torsion. If a torque is also applied to the beam, torsional shear stress must be calculated separately and superposed on 7 from Egs. 9.5.5. At a typical point H in the flange, Fig. 9.5.1, the largest principal stress has magnitude go; s+ z+ rather than simply o, as is commonly assumed. Usually, the discrepancy is small. A cross section of more than one material can be analyzed by using a substitute section that is homogeneous but has modified thickness. If the actual section is thin-walled, stresses are substantially constant through each thickness, and re- sistance to both bending and shear deformation is directly proportional to mod- ulus and to thickness. Accordingly, a modulus change in a thin-walled beam can be modeled by a thickness change. For example, if modulus £ in the web of an J section were twice the flange modulus E,, we could find q by analyzing a section with twice the actual web thickness but modulus £, throughout. How- ever, to find 7 in the web we must divide g by the actual web thickness. 2 “) +7? (compressive at point H in Fig. 9.5.1) (9-5-6) Unsymmetric Cross Sections. Two formulas for shear stress can be writ- ten. They differ in form but yield the same result. For each we resolve transverse shear force on the beam at the section in question into components V, and V.. Again we presume that the beam is straight, thin-walled, without taper, and homogeneous. ‘The foregoing derivation, which begins with Eq. 9.5.1, is little changed. It (a) (by tc) FIGURE 9.5.2. (a) Arbitrary cross section. Point O is its centroid, Loads V, and V, act through shear center S. (b) Various transverse shear loads (in the yz plane) on the end of a cantilever beam. Loads act through the shear center. (c) Shear flows that act on a tross section, Tinese flows are applied across section BB by the part of the beam nearest the wall. Views shown are in the negative x direction. Shear flows act on the hidden (rear) section BB. The dashed line is the neutral axis of bending. 9.5 SHEAR STRESS AND SHEAR FLOW IN STRAIGHT BEAMS: 363 is necessary only to use a @, expression appropriate to unsymmetric bending and to recognize that V, = dM,/dx and V, = —dM./dx, as shown in elementary texts. Directions of V, and V, in Fig. 9.5.2a, and directions of M, and M, in Fig. 9.2.2b, are consistent with the sign convention in the expressions V, = aM,/dx and V, = —dM,/dx. The first formula for shear stress is written by taking a, from Eq. 9.2.4. = Vly q _ (VA. + Vy + WV, 5. aaa (9.5.7) where = Ie dA and AZ = [ea (9.5.8) are first moments of the shaded area A, in Fig. 9.5.2a about the z and y axes, respectively, and and Z are coordinates of the centroid of area A,. Thus Eq. 9.5.7 gives 7 where thickness tis indicated in Fig. 9.5.2a. If y and z are principal centroidal axes, then /,. = 0, and Eq. 9.5.7 reduces to t Lt it which agrees with Eqs. 9.5.5. As is the case with Eqs. 9.5.5, Eqs. 9.5.7, and 9.5.10 below, presume that the cross section has a free edge, where g = 0. A closed cross section cannot be fully analyzed by these equations alone unless it has an axis of symmetry. Shear flow q will be zero on an axis of symmetry if the plane of load contains this axis. Cross sections that contain one or more closed cells are discussed in Section 9.9. The second formula for shear stress is written by taking o, from Eq. 9.3.2. q _ A = F tan dA) aaa ay (9.5.10) VAS 9 _ Ad (when L. = 0) (9.5.9) where angle A is defined by Eq. 9.2.5, and ¥ and 2 are again coordinates of the centroid of shaded area A, in Fig. 9.5.2a. Like Eq. 9.3.2, Eq. 9.5.10 requires that transverse loads on the beam share a common plane parallel to the x axis so that angle A of the neutral axis is constant along the beam. For the special case of principal axes, [,, = 0. If, in addition, V, = 0, then A = 0, and Eq. 9.5.10 reduces to the last term in Eq. 9.5.9. The algebraic sign of r in Eqs. 9.5.7, 9.5.9, and 9.5.10 may be misleading if used for assigning the direction of shear arrows on a cross section. To direct shear arrows properly, arguments of the following paragraph can be used. Directions of shear flows under various transverse loads are shown in Fig. 9.5.2c. Loads V, and V> are assumed to act parallel to principal axes of the cross section. Loads V, and V, act in the coordinate directions. In each of the four cases shown in Fig. 9.5.2c, the direction of g at any point can be established by arguments used in connection with Fig. 9.5.1. Further arguments are as follows. Under load V,, the direction of q changes at the bend. At the bend, 364 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS: shear flow is zero, as can be shown by symmetry considerations. Under load V2, there is no reversal of shear flow. Under load V,, the direction of q reverses as we travel from one end of the cross section to the other. The point of reversal is where the integral of 7, dA is zero (see Eq. 9.5.1). On passing this point, the algebraic sign of q changes. For equilibrium of forces in the yz plane, the integral of q dy along the horizontal leg must equal V,, and the integral of q dz along the vertical leg must vanish. Under load V,, similar comments apply. The reader is urged to sketch differential elements like the one in Fig. 9.5.1 so that the remarks of the foregoing paragraph are clearly understood. The flexure formula for o, is used to derive each expression for 7 in the present section. But the flexure formula is derived from the assumption that plane sections remain plane, which demands that r = 0. Despite this contra- diction, formulas for 7 are in good agreement with results of elasticity theory if the beam is slender, which in Fig. 9.5.1 means that L should be at least three or four times greater than b or h [2.1]. SHEAR CENTER: GENERAL REMARKS The nature of the problem is most easily grasped by thinking about a cantilever beam, Fig. 9.6.1. If loaded by a transverse force V, at its end, the beam will bend and probably twist as well. With different application points for load V,, the twist changes in magnitude and in sign. Clearly, there is some y-parallel line of action for V, that will produce zero twist. If a vertical load V, were applied, a corresponding z-parallel line for zero twist could be found. The two lines intersect at the shear center, also called the flexural center, which is the point through which a transverse end load of arbitrary inclination in the yz plane must pass if the beam is to bend without twisting. (Strictly, since in theory of elasticity an end load would be applied by distributed shear forces in the yz plane, we should regard an end load as a resultant force at the centroid of the distribution.) The foregoing argument admits different interpretations of the word ‘‘twist.’” FIGURE 9.6.1. Dashed lines show end de- flections of a cantilever beam for various loca- tions of a transverse end force V,. 8.6 SHEAR CENTER: GENERAL REMARKS 365 It might refer to rotation of the centroid, to average rotation of all points in the cross section, or to something else [9.2]. For a given shape of solid cross section, such as a triangle or a wedge, different definitions may lead to different locations of the shear center. The shear center of a solid or thick-walled cross section can be located by numerical methods [9.3]. Its position (by any plausible definition) is usually close to the centroid. The precise location is usually unimportant because a solid section does not twist greatly under a small torsional load. The location of the shear center is of much greater importance for a thin-walled open cross section because of its low torsional stiffness. The following example [9.4] illustrates concepts and shows that the shear center of an open cross section may not coincide with the centroid. If the tube of Fig. 9.6.2 is closed and V. acts at centroid O as shown, shear flow along line AB can be found from Eqs. 9.5.5. Since the vertical plane y = 0 is a plane of symmetry, g = 0 on this plane, and each lateral half of the beam is effectively an open section loaded by a transverse force V,/2. Along AB, (V./2)A,2 _ (V./2)@R1/2)2R/ 7) _ Vs rae aRt/2 aR (9.6.1) Now imagine that load V, is shifted a distance e to the left. With eccentricity e, load V. produces torque T = V.e about centroid O. Shear flow due to torque on the closed tube is, according to Eqs. 8.7.4, T = or (9.6.2) Along AB, this q is opposite in direction to the q of Eq. 9.6.1. What must be the distance e so that the resultant shear flow along AB is zero? We write vz, Ve mR = -2nR? = 0 therefore ¢ = 2R (9.6.3) But if the net shear flow is zero along AB, we can slit the tube open along AB without changing anything. Thus, if load V, acts through point S, Saint-Venant shear stresses associated with twisting of an open section (depicted in Fig. 8.4.3b) do not appear in Fig, 9.6.2 and therefore produce no twisting of the slit tube. We conclude that shear center S of the slit circular tube of Fig. 9.6.2 is at y = —2R. If loaded there, the very low torsional stiffness of the open section FIGURE 9.6.2. Thin-walled circular tube of constant thickness ¢. If closed, centrbid O isthe $ shear center. If slit open along AB, point S is the shear center. BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS: is not activated.' (To load the slit tube at point S, a bracket must be attached to its end.) The foregoing calculation method is simple for symmetric slit sections, but not for other open sections, such as channels and Z sections. Methods described in the following are more general. These methods locate the shear center by a moment equilibrium equation about an axis parallel to the beam axis. This equation requires that torque of the load balance torque produced by shear stresses on the cross section. The shear stresses are those associated with bending action and are described by equations such as Eqs. 9.5.5 or 9.5.7. Saint-Venant torsional shear stresses (Fig. 8.4.3) are not present and do not contribute to the moment equilibrium equation. With no twisting, relative axial motions along a slit (shown in Fig. 8.11.1a) do not arise. ‘A support may locally alter the position of the shear center. For example, in Fig. 9.6.2 the fixed support prevents axial displacements that prevail elsewhere in the cantilever beam. Stresses that enter the moment balance equation are modified near the fixed end. if the modified state of stress could be taken into account, it might be possible to locate load V, so that a given cross section along the beam does not rotate, although other cross sections might rotate small amounts, some clockwise and others counterclockwise. If distortion of the cross section in its own plane due to bending of the beam is neglected, the shear center coincides with the center of twist. (The center of twist is the point on the cross section that does not displace when a bar is twisted by a torque T.) That the two centers coincide can be proved by Maxwell’s reciprocal theorem, Imagine a cantilever, fixed at one end and loaded at the other. Let @, be rotation of the end cross section when loaded by an arbitrarily oriented transverse force V. Let A; be displacement of the point of application of load V, in the direction of V, caused by torque T. By the reciprocal theorem, VA, = Toy (9.6.4) If V is applied through the shear center, then 6 = 0. Therefore, Ay = 0, which is the definition of center of twist. Distortion of the cross section would inval- idate the argument, since T is applied by shear flows distributed over the cross section, and 6, would not be zero for all points in a distorted section. Even then the two centers should be close, if not coincident. Pioneer work in shear centers was done between 1913 and 1920, by Timo- shenko, by Griffith and Taylor, and especially by Maillart, the Swiss designer of graceful structures in reinforced concrete. "The shear center is not precisely ate = 2R. We argue as follows. Shear flow associated with torque T = Ve twists the tube clockwise at rate B = T/GJ = V.e/G2R%r. If V. acts a small distance de to the right ef the point y = —2R, Saint-Venant shear stresses appear. They are associated with the counterclockwise rate of twist 8 = AT/GJ, = V.de/(G2aR?'/3), calculated by Eqs. 8.4.8. The two rates of twist nullify one another if de = et?/3R. This calculation gives the shear center eccentricity as Pp ede = 2r(1 - 5) The correction Ac is negligible if R >> 1 9.7 SHEAR CENTER: OPEN SECTIONS—ONE AXIS OF SYMMETRY 367 9.7 SHEAR CENTER CALCULATIONS: OPEN SECTIONS WITH ONE AXIS OF SYMMETRY In this section and the next we consider straight beams that are homogeneous, thin-walled, without taper, and of open cross section. The calculation method described in the present section is less abstract than the method described in Section 9.8 but is straightforward only when the crass section has an axis of symmetry. Consider, for example, a channel section, like the beam of Fig. 9.6.1 but with a tip load V, acting parallel to the web (Fig. 9.7.1a). Horizontal distance e to shear center S is required, in terms of dimensions of the cross section. The method of calculation is easy to outline. If V. acts through S, there is no twisting and no Saint-Venant shear stress. The only forces parallel to the yz plane are those of Fig. 9.7.1¢, where F, and F,, are forces generated by shear flow q. For equilibrium, net moment about any point must vanish. (Strictly, we do not take moments about a point, but about an axis parallel to the longitudinal axis of the beam.) With point P an arbitrarily located moment center, Mp = V.e where = Mp = Fyh (9.7.1) The merit of locating point P on the web is that force F,, need not enter the calculations. All that remains is to express F; in terms of V, and dimensions of the cross section, Using Eqs. 9.5.5 to find q, at the corners in Fig. 9.7.1b, we have V.AZ _ Vi(bty(h/2) _ V.bhty ee eoteeeeteeer 2, (9.7.2) Since q varies linearly with distance across 4 flange, force F; is the average intensity of q times the width of the flange. Dt, @) 7 () © FIGURE 9.7.1. (a) Dimensions of a homogeneous channel section. Load V. is imagined as acting at the end of a cantilever beam. (b) Shear flows q that resist load V, (like the shear flows applied to the section of length / in Fig. 9.5.2b). (c) Resisting forces generated by shear flows q. 368 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS: Finally, calculating /, and using Eq. 9.7.1, we obtain 2 3 2, ee re) a L= 1p ft + 2f oe + m(*) | Ras + 7 (9.7.4) Mp _ Fh _ Bhs, 3b COU Ve a ht, + bby. BO (9.7.5) If the section is very deep or has narrow flanges, then b/h ~ 0 and e ~ 0, as is reasonable. The other limiting case is that of infinitely wide flanges, for which h/b = Oande = b/2. The z coordinate of the shear center of the foregoing channel section is zero. This conclusion is reached by considering a horizontal load V,. Shear flows become symmetrically distributed about the y axis, in both sign and magnitude. Flange forces F; act in the same direction and remain equal. Therefore, Mp = 0 in Eq. 9.7.1, and e = 0, where e is now the z coordinate of the shear center. In general, the shear center is always located on any axis of geometric and material symmetry if the cross section has such an axis. ko A similar agument jeads to the conclusion that a center of symmetry is also Hott a shear center. For example, point P is the center of symmetry of the homo- - geneous Z section in Fig. 9.7.2, as the geometry of either half is repeated by rotating the other half through 180° about point P. For any orientation of load V, arguments of Section 9.5 show that symmetrically located areas dA carry equal shear flows q. Resulting forces dF = q dA are equal and have zero net moment about point P. Thus Mp = 0 in Eq. 9.7.1, and distance ¢ from point P is zero. A simpler special case is that of an angle section, Fig. 9.7.3. Forces F, and F, are generated by shear flows q. These forces are concurrent at point P. Therefore, Mp = 0 in Eq. 9.7.1, and e = 0. Indeed, the same argument applies to unequal-leg angles and to bundles of rectangular plates that emanate from a single axis (Fig. 8.11.2): The point of concurrency is the shear center, whether or not the material is homogeneous. The shear center of a thin-walled section of more than one material can be located by analysis of an equivalent homogeneous section, as suggested in the paragraph that follows Eq. 9:5.6. ar ( e: e— y We oy B H iH P Hl =F be eat dF Equal legs Unequal legs? FIGURE 9.7.2. Homogeneous cross sec- FIGURE 9.7.3. The shear center of any tion with geometric symmetry about point P angle section is at the point of intersection of (which is therefore the centroid). the legs. 9.8 SHEAR CENTER: OPEN SECTIONS—NO AxIS OF SYMMETRY 369 9.8 SHEAR CENTER CALCULATIONS: OPEN SECTIONS WITH NO AXIS OF SYMMETRY Figure 9.8.1 shows a thin-walled, homogeneous open cross section of arbitrary shape. The beam is without taper. Axes y and z are centroidal axes of the cross section. They need not be principal axes. We ask for coordinates e, and e, of shear center S, where e, and e, are distances from point P, whose position is known. Point P can be located arbitrarily for convenience of calculation, as shown in a subsequent example. Before stating the main argument, we present a preliminary result that deals with A,y and A,Z in Eq. 9.5.7. Sectorial area w is defined and explained in Sections 8.9 and 8.10. We consider the integral of A,¥ dw over the entire cross section, from B to C in Fig. 9.8.1. Integrating by parts, we have c Jasao = Ajo} — [eam = - | yoda (9.8.1) B Term A,¥e |§ vanishes because at end points (B and C) A,¥ is the first moment of the entire cross-sectional area about the z axis. This moment is zero because z is a centroidal axis. In the remaining integral, d(A,) = y dA by Eqs. 9.5.8. The integral of Az dw is transformed similarly. Collecting results and assigning the names S,,, and S,, to the integrals, we have 4 | yoda = -[ Ay de S. -{ wwda = ~[ Ado area area (9.8.2) 5. “y where S,, and S,, are recognized as sectorial linear moments, defined in Eqs. 8.10.1. In Fig. 9.8.1 an arbitrarily oriented force, whose components are V, and V., is assumed to act through shear center S. Resisting shear flow g is generated by V, and V, and acts on the rear face of the section, in the manner of Fig. 9.5.2. For no twisting of the beam, net moment about any point must vanish. Choosing point P, we write -Vye. + Vey + Sgrds = 0 (9.8.3) FIGURE 9.8.1. | Thin-walled open cross section of arbitrary shape. Flow q is the resist- ing shear flow generated by loads V, and V,. Axes yz are centroidal. 370 BENDING, SHEAR CENTER, 4NO THIN-WALLED BEAMS: Now rds = dw, and q is given by Eq. 9.5.7. Thus, with the symbols of Eqs. 9.8.2, Eq. 9.8.3 becomes + v( " 1 Sey — 1 V,( -e, - This equation can be true for arbitrary values of V, and V, only if the paren- thetica] expressions vanish separately. Therefore, Sun — 1y:Suy Sux — bSuy Pca eid eas OS hi = Th =0 (9.8.4) (9.8.5) These equations for e, and e, locate the shear center of a homogeneous, thin- walled open cross section of arbitrary shape. The choice of initial point for the sectorial area diagram is arbitrary. A different choice changes w by a constant but leaves S,,, and S,,, unchanged. Example 9.8.1. We illustrate the use of Eqs. 9.8.5 by considering a section with one axis of symmetry, Fig. 9.8.2. Thus calculations are simple and results are easily checked by applying arguments of Section 9.7. With pole P at one corner and initial point / at another, sectorial area w is Zero over all but one Jeg of the channel. By the method of Eq. 8.10.8, with w the general function, —4a? | 2a 8a't ot eee ae set 0.8.6) Se Qa)(a) = ~4a't Applying Eqs. 9.8.5, with /’s as given in Fig. 9.8.2, we find —4a't _ ba 14a°1/3 7 (9.8.7) w= 4a? FIGURE 9.8.2. Calculation of the shear center location for a thin-walled channel section. Axes y and z are centroidal. 8.8 GHEAR AND SHEAR CENTER IN CLOSED SECTIONS 371 That e, is negative means that S is to the left of P. Equations 9.8.7 indicate that shear center S is on the y axis a distance 6a/7 left of the web. (It is a worthwhile exercise to check that the same position of S relative to the channel is obtained if the section is rotated 180° about point O so that the web crosses the +y axis.) SHEAR AND SHEAR CENTER IN CLOSED SECTIONS We consider thin-walled straight beams without taper that contain one or more closed cells. Torsion of single and multicell tubes is considered in Chapter 8. Here we consider effects of transverse loads. Under the assumption that plane sections remain plane, flexural stress @, can be found by Eq. 9.2.4 or by Eq. 9.3.2. Finding shear flows and the shear center is more troublesome. We cannot simply use Eq. 9.5.7, or shear center calcu- lations that follow from it, because Eq. 9.5.7 requires having a free edge (ar knowing some other point at which q vanishes), as a review of the derivation will show. We do not know in advance where q = 0 for a general closed section. The analysis procedure is to introduce free edges by cutting cells open, then close them again by finding the shear flow in each cell that makes all cells have the same rate of twist. Consider, for example, the two-cell, thin-walled cross section of Fig. 9.9.1. Load V,, its position y, and geometry of the section are regarded as known quantities. Shear flows and rate of twist 8 are required. Shear flows in Fig. 9.9.1b, which are present after both cells are cut open, are arbitrarily assumed to act counterclockwise about arbitrary reference point P. These shear flows can be found by applying Eq. 9.5.7. Shear flows q., and 4-2 teleased by the cut are as yet unknown. (Note that flows q,, and q,. are constant around each cell, whereas flows g,,, 9,2, and q,,,. vary.) Net shear flows are NM = Gor + Ger 92 = Ir + Ger (9.9.1) Iw = Gow + Ger ~ Ger Three equations are needed to solve for the unknowns, which are q.,, q.2, and B. Two equations come from Eq. 8.8.2. They say that each cell twists the same amount. Algebraic signs are chosen such that q., and q.2 make positive contri- butions to B along all paths in their respective cells. The two twist equations are ! a | we B 2GI, (x t a Nae t a etl gs [te ) 8 war, (loc ## cat as (9.9.2) PROBLEMS 395 9.3. W. E. Mason and L. R. HERRMANN, “Elastic Shear Analysis of General Prismatic Beams,” Journal of the Engineering Mechanics Division (Proc. ASCE), Vol. 94, No. EM4, 1968, pp. 965-983. 9.4. M. J. French, “*A Simple Introduction to Shear Centers," International Journal of Mechanical Engineering Education, Vol. 2, No. 2, 1974, pp. 55-57. PROBLEMS* Section 9.1 91 Prove the first three transfer theorems (Eqs. 9.1.2). 9.2 (a) Derive Eqs. 9.1.5 from Eqs. 9.1.4. (b) Derive Eq. 9.1.6. (c) Derive Eq. 9.1.7. *9.3 (a) For the Z section shown, calculate /,, /,, and /,,. Also locate the principal centroidal axes and calculate the principal moments of inertia. (b) Calculate /,, /,, and /,,, for the quarter circle. Also find [nox and [jyin Teferred to centroidal axes. Express answers in terms of R. Section 9.2 9.4 95 9.6 *9.7 9.8 Find expressions for b and c in Eq. 9.2.1, and verify Eqs. 9.2.4 and 9.2.5. Derive Eq. 9.3.2 from Eqs. 9.2.4 and 9.2.5. Show that for a circular cross section of outer radius R, Eq. 9.2.6 yields G, = MR/I, where M = VM? + M? and / is taken about a diameter. re max (a) In Fig. 9.2.2, let 8 = 0, b = 30 mm, and h = 50 mm. Find the orientation of the neutral axis. Also calculate the largest tensile and compressive flex- ural stresses in terms of the applied bending moment M. (b) Show that if 8 = 0 in Fig. 9.2.2, the neutral axis always passes through corner B of the right triangle, regardless of the values of b and h. A force P = 8000 N acts upward and to the left on the end of a 750-mm-long cantilever beam that has the triangular cross section shown. At the fixed end, | wet bs 20 nin ——}S0 mm. : mm | pe 30°F | ‘| 7 ceuele, il YY) atl <7" PROBLEM 9.3 ; PROBLEM 9.8 20mm 80mm 100mm *An asterisk before a problem number indicates that a full or partial answer is given in the back of the book 9.9 9.10 9 9.12 *9,13 *9.14 *9,15 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS compute flexural stress at corners A, B, and C. Also focate the neutral axis with respect to the y axis. What centroidal axial force must be applied to the beam of Fig. 9.2.3 to make lo.4| = |a,g| after bending and axial stresses are superposed? Repeat Example 9.2.2 with reference to the I section of Problem 9.15. Flexural stress o;, in the rectangular cross section shown can be calculated using either the yz or the & coordinate system. Given that J, = —53.76 mm* and that moment M, is the only load, show that the two methods give the same result at point A. Repeat for point B. A cantilever beam has a rectangular cross section. The plane of loads passes through diagonally opposite corners of the cross section. Show that the other two corners are free of flexural stress. (a) A cantilever beam 1.6 m long has the Z section shown. It carries a load P = 2500 N in the —z direction at the free end. Find the maximum magnitude of flexural stress (either tensile or compressive). Where does this stress appear? (b) Repeat part (a) with load P redirected so that it acts in the —y direction. f6 mm— 10-4 boss 55 10 PROBLEM 9.11 PROBLEM 9.13 Imagine that a simply supported beam, 3 m long, has the quarter-circle cross section shown in Problem 9.3(b). A uniformly distributed downward load (in the —z direction) of 3500 N/m is applied. What must R be if the tensile flexural stress is not to exceed 24 MPa? A column has the I section shown. A compressive force P parallel to the axis lee J, = 1105 (10)3 mmé 100 mm —y 1, = 49.410) mm4 A = 680 mm? D c Fem PROBLEM 9.15 PROBLEMS 397 of the column is applied at corner A. Find axial stress o, in terms of P at the remaining three corners B, C, and D. 9.16 Repeat Problem 9.15, but let axial load P be applied to one corner of a rectan- gular cross section of dimensions b by h. In addition, sketch the distribution of a, over the cross section in the manner of Fig. 9.2.2 9.17 A column with a rectangular cross section of dimensions b by h carries an axial compressive force. If the force acts within a certain central area of the cross section, there will be no tensile stress in the column. Define this area, known as the kern, in terms of 6 and / and show it on a sketch. Section 9.3 9.18 (a) In Example 9.3.1 let q = SQ/L. Calculate the magnitude and direction of the resultant end deflection A in terms of Q, L, and E. (b) Show that the same result is given by working with load components parallel to principal axes of the cross section, as suggested early in Section 9.3. *9,19 A cantilever beam has a narrow rectangular cross section, 40 mm by 500 mm. The axis of the beam is horizontal, but the 500-mm dimension is inclined at 6° to the vertical. The beam deflects under its own weight. Compute the angle with respect to the horizontal of the end deflection. 9.20 A short column has the angle cross section depicted in Fig. 9.1.2. A horizontal bar is attached to the top at an angle a@ with respect to the short leg of the angle, as shown. Under load P, parallel to the column and 1 m from its centroidal axis, deflection of the top of the column is to have no component parallel to the longer leg of the cross section. What should @ be? If P = 400 N and E = 200 GPa, what is the deflection of the top of the column? *9,21 The cross section shown is that of a simply supported beam 2 m long, loaded by a uniformly distributed load g whose plane is 20° from the vertical. Compute the magnitude and direction of the center deflection in terms of q and elastic modulus E. (a) By the method of Eqs. 9.3.3 and 9.3.4. (b) By using components referred to principal axes. 9.22 Compute the magnitude and direction of the deflection due to bending at the end of the cantilever beam shown. The cross section is an unequal-leg channel. Let E = 70 GPa. ad i 2.0m ae 4 710) mm th, Jy = 0.410) mm4 ys PROBLEM 9.20 PROBLEM 9.21 PROBLEM 9.22 398 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS Section 9.4 9.23 (a) Verify that Eqs. 9.4.4 yield the expected result in the limiting case of zero pretwist. (b) Show that F,, F,, 4,, and A, satisfy Maxwell's reciprocal theorem. (c) A pretwisted bar has a narrow rectangular cross section. Show that the loading discussed in Section 9.4 does not cause the bar to twist. (a) Imagine that tip forces F, and F. in Fig. 9.4.1 are replaced by tip moments M, and M, whose vectors point in the +y and +z directions, respectively Revise Eqs. 9.4.1 through 9.4.3 as required if tip deflections are again to be computed. *9.24 A strip of metal with a | mm by 10 mm rectangular cross section is pretwisted a full 180° between ends (see the sketch). It is to serve as a tip-loaded cantilever beam 120 mm long. (a) Apply a tip load P in the y direction. Compute the y and z components of tip deflection in terms of P and E. Explain why the signs and relative magnitudes of A, and A. are reasonable or unreasonable (b) Repeat part (a) for a tip load P in the z direction (c) What would be the deflections in parts (a) and (b) if the beam were not pretwisted? In comparison with deflections of the pretwisted beam, why are these results reasonable or unreasonable? (d} Compute the maximum flexural stress in part (a) in terms of P. ies eee a PROBLEM 9.24 Section 9.5 9.25 In the T section of Fig. 9.5.1, let Q = 1000 N, L = 400 mm, h = 80 mm, and b = 70 mim. Let the flange have constant thickness 6 mm and the stem have constant thickness 4 mm. (a) Show by sketches how shear flow q and shear stress 7 vary across the flange and along the stem. Compute numerical values at several stations (b) Evaluate the largest normal stress in the flange from Eq. 9. percentage is this stress larger than a, at the same location? , 9.26 A cantilever beam in the form of a slit cylindrical tube is loaded by a force V that passes through its shear center so that the beam bends without twisting (see PROBLEMS 399 *9.27 9.28 9.29 *9.30 v R Af la Ss y v PROBLEM 3.26 the sketch). Thickness 7 is constant, and R >> ¢. Find the values of a for which the shear flow is zero, and compute the maximum magnitude of shear flow if (a) b=0. (b) b= w/2. (©) Wis an arbitrary angle. For the equal-leg angle in Fig. 9.5.2b, let h = 90 mm and let the thickness be 10 mm throughout. Also let each load have unit magnitude. (a) Calculate the location and magnitude of the maximum shear stress produced by load V,. Use principal axes of the cross section (b) Repeat part (a) for load V. (c) For load V,, locate where 7 reverses direction. Also compute the location and magnitude of the largest 7. Use Eq. 9.5.7. Repeat Problem 9.27, using Eq. 9.5.10 throughout. However, in part (c) use V, instead of V,. (a) In Problem 9.27, let V, = 0.707 N and V, = —0.707 N. By superposing results, show that when V, and V, act simultaneously there is zero shear Stress at the bend (as is given by load V, acting alone). In Problem 9.27, let V, = V, = 0.707 N. By similar superposition, show that when V, and V, act simultaneously, shear stress at the bend is the same as given by load V, = 1.0 N acting alone. (a) Let a y-parallel shear force V, act on the Z section of Problem 9.13. In the z-parallel leg 0 < z < 70 mm, find an expression for shear flow q in terms of V,. Also find the maximum value of q and where it appears in this leg. (b) Repeat part (a), but consider the y-parallel leg 0 < y < 55 mm. (ce) Sketch the variation of q over the entire cross section. At what locations does q vanish? (d) One might ignore the asymmetry and make the estimate = (V,/Ayendt, as is commonly done for an I section. What is the percentage error of this estimate? () Section 9.6 *9.31 Use the calculation method of Section 9.6 to locate the shear center of a slit thin-walled rectangular tube whose cross section is shown. Thickness ¢ is con- stant. Imagine that the tube is slit lengthwise at (a) Midside B. (b) Corner D. 400 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS PROBLEM 9.31 Section 9.7 9,32, *9.33, 9.34, 9.35, *9.36, 9.37-9.40, *9.41, *9.42, 9.43 The open cross sections shown are all thin-walled, of constant thickness, and have one axis of symmetry. Assume a frictionless contact where overlaps appear. In each case find an expression for the location of the shear center. Check your answer by examining limiting cases where possible (as by letting dimension b approach zero, for example). PROBLEM 9.32 ig 7 © a ek PROBLEM 9.35 PROBLEM 9.37 — ae 4+ jiniiisttteaa ei PROBLEM 9.38 PROBLEM 9.39 PROBLEM 9.40 PROBLEMS 401 SF | ale ¢ Le ’ PROBLEM 9.41 PROBLEM 9.42 PROBLEM 9.43 9.44 (a) Several beams of arbitrary cross section have their centroids in a common plane AB. If somehow coupled together so that they share the same neutral axis, the composite beam will bend without twisting if ¢ is as stated in the sketch. Derive this expression for e, in which y; is the distance to the ith shear center. (b) Use this formula to solve Problem 9.35. (c) Use this formula to solve Problem 9.43. Eyl, Egly Ext; Egha Exly Ely, t Eyhyys tt Elva sty QYB--Gon Ba ee mje] al 2 a y PROBLEM 9.44 Section 9.8 9.45 For an arbitrary open cross section, prove that S,, = S,, = 0 if axes yz are centroidal and pole P coincides with shear center §, as claimed above Eq. 8.10.5. Suggestion: Sete, = e, = 0 in Eqs. 9.8.5. 9.46 Use Egs. 9.8.5 to show that point P is the shear center in Figs. 9.7.2 and 9.7.3. 9.47 Use the method of Section 9.8 to locate the shear centers of the following cross sections. (a) Problem 9.35. (b) Problem 9.36. (c) Problem 9.40 (d) Problem 9.42. 9.48 — Imagine that the box section of Problem 9.31 is cut open at the lower left corner. Locate the shear center by the method of Section 9.8. 9.49 Locate the y and z coordinates of the sheat center of the cross section shown. Dimensions are in millimeters. Thickness 1 is constant. Axes yz centroidal, and 1, = 10.5(10)° mm‘, /,, = 20.8(10)° mm‘, /,, = 6.00(10)° mm‘. *9.50 Imagine that the channel section of Fig. 9.7.1 is thin-walled, of constant thick- ness, with h = 5, but that the upper flange has a modulus three times as great as that of the rest of the section. Locate the shear center. 402 BENDING, SHEAR CENTER, AND THIN-WALLED BEAMS 4 #0 iz PROBLEM 9.49 PROBLEM 9.51 PROBLEM 9.52 Section 9.9 9.51 Find expressions for rate of twist and maximum shear stress in the thin-walled, homogeneous closed tube of triangular cross section when loaded as a cantilever by end load V, (see the sketch). Neglect stress concentrations. *9,52 Locate the shear center of the thin-walled, nonhomogeneous closed D section shown. Also determine the maximum shear stress caused by a unit vertical load through the shear center. Neglect stress concentrations. 9,53 The outer skin of the three-cell box section shown has thickness 1.0 mm. The two internal webs have thickness 0.5 mm. After applying a unit vertical load and cutting the cells open at A, B, and C, web shear forces shown in the second sketch are found. Locate the shear center of the closed three-cell section. a 20 mm 4 oe ae 0.100N 0.225N 0.400N PROBLEM 3.53 Section 9.10 9.54 Derive Eq. 9.10.4 from Eqs. 9.10.2 and 9.10.3. 9.55 Can axial forces apply a bimoment to the cross sections depicted in Fig. 8.11.2? Explain. *9.56 Imagine that a bimoment equal to the bimoment Fbh in Fig. 9.10.2 is to be applied by axial stress o, distributed over the flanges at x = L. Describe the magnitude and variation of o, atx = L if (a) The variation is linear, 0, = *+ky (b) a; is constant over each half-flange b/2; o, = *constant. *9.57 Axial forces F are applied to the ends of the Z section of Fig. 9.10.3 as shown. Additional axial forces are to be applied to the ends to keep the beam from

You might also like