Skip to main content
Log in

Thermoelastic Determination of Individual Stresses in Vicinity of a Near-Edge Hole Beneath a Concentrated Load

  • Published:
Experimental Mechanics Aims and scope Submit manuscript

Abstract

Thermoelastic data are combined with an Airy stress function to determine the individual stresses on and near the boundary of a circular hole which is located below a concentrated edge-load in a plate. Coefficients of the stress function are evaluated from the measured temperatures and the local traction-free conditions are satisfied by imposing \( {\sigma_{r{\rm{r}}}} = {\tau_{r\theta }} = 0 \) analytically on the edge of the hole. The latter has the advantage of reducing the number of coefficients in the stress function series. The method simultaneously smoothes the measured input data, satisfies the local boundary conditions and evaluates individual stresses on, and in the neighbourhood of, the edge of the hole. Attention is paid to how many coefficients to retain in the stress function series. Although the presence of high stress concentration factors, together with a hole-diameter-to-plate-thickness ratio of only two, result in some three-dimensional effects, these are relatively small and the agreement between the thermoelastic values, those from recorded strains and FEM-predicted surface stresses is good.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. The combined photoelastic stress function technique of Ref. 5 is effective for overcoming the difficulties associated with time-edge affects that can adversely affect photoelastic stress analyses and which might be present in any of the isochromatic fringe patterns of Ref. 2.

  2. These evaluations could be supplemented by also imposing σ  = 0 of equation (A3) at points on the edge of the hole, but this latter action/equation cannot by itself evaluate all of the coefficients.

Abbreviations

a n , b n c n and d n :

Airy coefficients

A:

Airy matrix

c :

vector of Airy coefficients

C :

condition number

C 1 . . . . , , 9 :

coefficients

d :

thermoleastic S input vector

d′ :

vector of evaluated S

D :

location of hole below top edge of plate

k :

measure of number of Airy coefficients

K :

thermo-mechanical coefficient

m :

number of input values

P :

load per thickness

P* :

concentrated load

r, θ :

polar coordinates

R :

radius of hole

σ:

isopachic stress \( ( = {\sigma_{rr}} + {\sigma_{\theta \theta }}) \)

S* :

thermoelastic recorded signal

t :

thickness of plate

Δ:

change

ϕ :

Airy stress function

σ :

normal stress

τ :

shear stress

References

  1. Greene RJ, Patterson EA, Rowlands RE (2008) Thermoelastic stress analysis, Ch. 26, Handbook of Experimental Solid Mechanics, ed. by W. Sharpe, Springer

  2. Wang W-C, Chen M-Y, Lin M-S, Wu C-P (2005) Investigation of the stress field of a near surface circular hole. Exp Mech 45(3):244–249

    Article  Google Scholar 

  3. Lin, S-J. (2007) Two- and three-dimensional hybrid photomechanical-numerical stress analysis, Ph.D. Thesis, University of Wisconsin, Madison, WI, USA

  4. Lin SJ, Matthys DR, Rowlands RE (2009) Individual stresses in perforated plates by thermoelasticitiy and Airy stress function. Strain 45:511–526

    Article  Google Scholar 

  5. Foust BE (2003) Individual stress determination in Inverse problems by combining experimental methods and Airy stress functions, Thesis, University of Wisconsin, Madison, WI, USA

  6. Foust BE, Rowlands RE (2003) Inverse stress analysis of pinned connections using strain gages, Inverse Problems in Engineering Mechanics (ISIP 2003), Nagano, Japan, M. Tanaka, editor: 323–332

  7. Ryall TG, Heller M, Jones R (1992) Determination of stress components from thermoelastic data without boundary conditions. J Appl Mech 59:841–847

    Article  Google Scholar 

  8. Ryall TG, Cox PM, Enke NF (1992) On the determination of dynamic and static stress components from measured thermoelastic data. Mech Mater 14:47–57

    Article  Google Scholar 

  9. Joglekar N (2009) Separating stresses using Airy’s stress function and TSA: Effects of varying the amount and source locations of the input measured TSA data and the number of Airy coefficients, Thesis, University of Wisconsin, Madison, WI, USA

  10. Press WH, Flannery BP, Teukolsky SA, Vetterling WT (1996) Solution of Linear Algebraic Equations, Chapter 2, “Numerical Recipes in C”, 2nd Edition, Cambridge University Press

  11. Schilling RJ, Harris SL (2000) Applied numerical methods for engineers using MATLAB and C, Chapter 2, “Linear Algebraic Systems”. Brooks/Cole, New York

    Google Scholar 

  12. Dulieu-Barton JM, Stanley P (1998) Development and applications of thermoelastic stress analysis. J Strain Anal 33(2):93–104

    Article  Google Scholar 

  13. Huang YM, Rowlands RE (1991) Quantative stress analysis based on the measured trace of the stress tensor. J Strain Anal 26(1):55–63

    Article  Google Scholar 

  14. Lin S-T, Rowlands RE (1995) Thermoelastic stress analysis of orthotropic composites. Exp Mech 35(3):257–265

    Article  Google Scholar 

  15. Lekhnitskii SK (1968) Anisotropic plates. Gordon Breach Science, New York

    Google Scholar 

  16. De Jong T (1981) Stresses around rectangular holes in orthotropic plates. J Compos Mater 15:311–327

    Google Scholar 

Download references

Acknowledgements

The authors wish to thank the US Air Force for providing funds (Grant #FOSR FA9550-05-1-0289) with which to purchase the TSA equipment, and John Dreger and Weston Skye, University of Wisconsin, for their assistance. Dr. Quinn’s participation was funded by the Worldwide Universities Network (WUN), which enabled him to visit the University of Wisconsin-Madison under the Global Exchange Programme. Dr. Lin acknowledges the kind support of the Robert M. Bolz Wisconsin Distinguished Graduate Fellowship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. E. Rowlands.

Appendices

Appendix A: Equations

From equations (1) and (2), the stresses at any point in a loaded plate such as that of Fig. 1 can be written in terms of polar coordinates r, θ (coordinate origin at distance D below the top edge of the plate) as follows [3]

$$ \begin{gathered} \,\,\,\,{\sigma_{rr}} = \left\{ \begin{gathered} \frac{{2 \cdot \left( {D - r \cdot \cos \theta } \right)}}{{\pi \cdot {{\left( {{r^2} + {D^2} - 2 \cdot D \cdot r \cdot \cos \theta } \right)}^2}}} \cdot \hfill \\\left[ \begin{gathered} - {\left( {D - r \cdot \cos \theta } \right)^2} \cdot {\cos^2}\theta - {r^2} \cdot {\sin^4}\theta \hfill \\+ r \cdot \sin \theta \cdot \left( {D - r \cdot \cos \theta } \right) \cdot \sin \left( {2 \cdot \theta } \right) \hfill \\\end{gathered} \right] \hfill \\\end{gathered} \right\} \cdot P \hfill \\+ \frac{{{b_0}}}{{{r^2}}} + 2 \cdot {c_0} + \left( { - \frac{{2 \cdot {c_1}}}{{{r^3}}} + 2 \cdot {d_1} \cdot r} \right) \cdot \cos \theta \hfill \\- \sum\limits_{n = 2,3,4...}^N {\left[ \begin{gathered} {a_n} \cdot n \cdot \left( {n - 1} \right) \cdot {r^{\left( {n - 2} \right)}} + {b_n} \cdot \left( {n + 1} \right) \cdot \left( {n - 2} \right) \cdot {r^n} \hfill \\+ {c_n} \cdot n \cdot \left( {n + 1} \right) \cdot {r^{ - \left( {n + 2} \right)}} + {d_n} \cdot \left( {n - 1} \right) \cdot \left( {n + 2} \right) \cdot {r^{ - n}} \hfill \\\end{gathered} \right] \cdot \cos \left( {n \cdot \theta } \right)} \hfill \\\end{gathered} $$
(A1)
$$ \begin{gathered} \,\,\,{\sigma_{\theta \theta }} = \left\{ \begin{gathered} \frac{{2 \cdot \left( {D - r \cdot \cos \theta } \right)}}{{\pi \cdot {{\left( {{r^2} + {D^2} - 2 \cdot D \cdot r \cdot \cos \theta } \right)}^2}}} \cdot \hfill \\\left[ { - {{\left( {D - r \cdot \cos \theta } \right)}^2} \cdot {{\sin }^2}\theta - {r^2} \cdot {{\sin }^2}\theta \cdot {{\cos }^2}\theta - r \cdot \sin \theta \cdot \left( {D - r \cdot \cos \theta } \right) \cdot \sin \left( {2 \cdot \theta } \right)} \right] \hfill \\\end{gathered} \right\} \cdot P \hfill \\- \frac{{{b_0}}}{{{r^2}}} + 2 \cdot {c_0} + \left( {\frac{{2 \cdot {c_1}}}{{{r^3}}} + 6 \cdot {d_1} \cdot r} \right) \cdot \cos \theta \hfill \\+ \sum\limits_{n = 2,3,4...}^N {\left[ \begin{gathered} {a_n} \cdot n \cdot \left( {n - 1} \right) \cdot {r^{\left( {n - 2} \right)}} + {b_n} \cdot \left( {n + 1} \right) \cdot \left( {n + 2} \right) \cdot {r^n} \hfill \\+ {c_n} \cdot n \cdot \left( {n + 1} \right) \cdot {r^{ - \left( {n + 2} \right)}} + {d_n} \cdot \left( {n - 1} \right) \cdot \left( {n - 2} \right) \cdot {r^{ - n}} \hfill \\\end{gathered} \right] \cdot \cos \left( {n \cdot \theta } \right)} \hfill \\\end{gathered} $$
(A2)
$$ \begin{gathered} \,\,\,\,{\tau_{r\theta }} = \left\{ \begin{gathered} \frac{{2 \cdot \left( {D - r \cdot \cos \theta } \right)}}{{\pi \cdot {{\left( {{r^2} + {D^2} - 2 \cdot D \cdot r \cdot \cos \theta } \right)}^2}}} \cdot \hfill \\\left[ \begin{gathered} \frac{1}{2} \cdot {\left( {D - r \cdot \cos \theta } \right)^2} \cdot \sin \left( {2 \cdot \theta } \right) - \frac{1}{2} \cdot {r^2} \cdot {\sin^2}\theta \cdot \sin \left( {2 \cdot \theta } \right) \hfill \\+ r \cdot \sin \theta \cdot \left( {D - r \cdot \cos \theta } \right) \cdot \cos \left( {2 \cdot \theta } \right) \hfill \\\end{gathered} \right] \hfill \\\end{gathered} \right\} \cdot P \hfill \\+ \left( { - \frac{{2 \cdot {c_1}}}{{{r^3}}} + 2 \cdot {d_1} \cdot r} \right) \cdot \sin \theta \hfill \\+ \sum\limits_{n = 2,3,4...}^N {\left[ \begin{gathered} {a_n} \cdot n \cdot \left( {n - 1} \right) \cdot {r^{\left( {n - 2} \right)}} + {b_n} \cdot n \cdot \left( {n + 1} \right) \cdot {r^n} \hfill \\- {c_n} \cdot n \cdot \left( {n + 1} \right) \cdot {r^{ - \left( {n + 2} \right)}} - {d_n} \cdot n \cdot \left( {n - 1} \right) \cdot {r^{ - n}} \hfill \\\end{gathered} \right] \cdot \sin \left( {n \cdot \theta } \right)} \hfill \\\end{gathered} $$
(A3)

As indicated in section 2, “Relevant Equations,” the stress function of equation (1), and hence the stresses of equations (A1) through (A3), assume symmetry about the x-axis; single-valued stresses, strains and displacements; and the loaded member can contain a hole at distance D below the top of the plate of Fig. 1. However, the shape of the hole is not yet defined. While ϕ of equation (1) contains \( {a_0} \) and \( {a_1} \), these coefficients disappear during the differentiations of equations (2) and hence do not occur in any of the stresses. Individual stresses could be obtained from equations (A1) through (A3) if the Airy coefficients, b 0 , c 0 , c 1 , d 1 , a n , b n , c n and d n , and P were known, and without knowledge of constitutive information (other than assuming elastic isotropy), geometry, or other external boundary conditions. For load P, and circular hole of radius R and centered at location D below the top of the plate, values of the coefficients b 0 , c 0 , c 1 , d 1 , a n , b n , c n and d n of equations (A1) and (A2) could be obtained by imposing σ rr  = 0 of equation (A1) at r = R at discrete locations around the holeFootnote 2. Having determined these Airy coefficients based on σ rr  = 0 at the edge of the hole, the hoop and shear stresses, σ θθ, and σ , would then be available from equations (A2) and (A3). However, without even addressing how many coefficients to retain and/or at how many positions to impose this traction-free condition at r = R, it is unlikely the resulting components of stress would be reliable except at positions very close to the edge of the hole. On the other hand, combining the imposed tractions-free conditions at r = R with TSA-measured values of σ = σ rr  + σ θθ away from the hole enables the three individual stresses of equations (A1) through (A3) to be evaluated reliably at and in the vicinity of the hole. Since neither b o , c 1 , a n , nor c n would appear in the isopachic stress σ = σ rr  + σ θθ based on equations (A1) and (A2), these Airy coefficients cannot be determined solely from experimentally-determined values of the isopachic stress, σ, based on these expressions, thereby necessitating combining some local collocation or other information with measured data.

As noted previously, equations (A1) through (A3) contain coefficients that do not appear in the isopachic stress σ = σ rr + σ θθ based on equations (A1) and (A2), thereby precluding determining the individual stresses of equations (A1) through (A3) from equations (A1) and (A2) and TSA alone. However, imposing the traction-free conditions analytically at r = R for all θ (this also defines the shape of the hole) results in identical coefficients existing in the isopachic stress expression (σ = σ rr + σ θθ) and individual stresses of equations (4) through (16) [3]. In addition to now being able to evaluate the remaining Airy coefficients from only TSA data, incorporating these local boundary conditions at the hole analytically enables one to acquire accurate stresses with fewer coefficients (imposing such boundary conditions analytically typically reduces the number of independent Airy coefficients by about 50% compared with doing so non-analytically/discretely) and consequently potentially needing fewer measured input data [3].

Appendix B: Three-Dimensional FEM

Acknowledging the two planes of symmetry of the plate of Fig. 1, a quarter symmetry 3-D ANSYS model was used. To prevent rigid body motion, nodes on the bottom surface of this plate were constrained to have zero displacements in the vertical x-direction. Post-processing checks confirmed that reaction forces at all these nodes were in the positive x-direction. The applied load P of Fig. 1 was simulated by constraining those nodes on the top surface of the plate which lay in the symmetry plane containing the axis of the hole to have a small fixed displacement (0.1 mm) in the negative x-direction. The reaction force acting at these nodes was determined in a post-processing step and used to scale the resulting stresses to the desired applied force (889.6 N = 200 pounds), and then those stresses were normalized using the uniform stress of 1.05 MPa (152.38 psi). This approach avoids the need to assume a force distribution on the top surface of the plate and gives a good approximation of the experimental application of load using a roller.

Three-dimensional mesh density was varied by controlling the number of element divisions around the circumference of the hole. Element sizes in the remainder of the model were linked to this parameter so that a change in mesh density around the hole propagated in a controlled way throughout the model. Models with 10, 20 and 40 element divisions around one quarter of the hole circumference were analyzed, see Fig. 14. Linear hexahedral brick elements (ANSYS type SOLID45) were used. Some degenerate (prism) elements were generated in the meshing process, but these were all well away from the regions of interest, the latter being above and around the hole boundary where approximately equilateral elements were ensured by using mapped meshing.

Fig. 14
figure 14

Quarter models of the plate for number of elements NEL = (a) 10, (b) 20 and (c) 40, respectively

The stability of the finite element solution as a function of mesh density was assessed by plotting the stresses on the boundary of the hole as a function of mesh density, see Figs. 15 and 16. Even for the coarsest mesh, the hoop stress at the plate surface on the hole boundary at the side of the hole, θ = 0 of Figs. 15 and 16, is 97.5% of the value for the finest mesh. This suggests that the difference between the solution for the finest mesh and the “true” solution is small. All subsequent results were generated using this finest mesh of Fig. 14(c), NEL (number of elements/mesh density) = 40. Component stresses on the external surface at a radius r = 1.17 R are plotted in Figs. 17 and 18, whereas Figs. 19, 20 and 21 illustrate these stresses at r = 1.17 R also as a function of angular position, θ, but through the plate thickness (center of plate at left in these figures).

Fig. 15
figure 15

Three-dimensional ANSYS normalized hoop stress, σ θθ / σ o , at the hole boundary as a function of θ on the external surface of the plate for the various mesh densities of Fig. 14

Fig. 16
figure 16

Three-dimensional ANSYS normalized stresses on the external surface of the plate at θ = 0 at the hole boundary for various mesh densities of Fig. 14

Fig. 17
figure 17

Three-dimensional ANSYS normalized circumferential stress, σ θθ / σ o , at r = 1.17R on the external surface of the plate for a mesh density of NEL = 40

Fig. 18
figure 18

Three-dimensional ANSYS normalized shear stress τ / σ o , at r = 1.17R on external surface of the plate for a mesh density of NEL = 40

Fig. 19
figure 19

Three-dimensional ANSYS normalized circumferential stress, σ θθ/ σ o , at r = 1.17R through the plate thickness and at various angular positions θ for a mesh density of NEL = 40

Fig. 20
figure 20

Three-dimensional ANSYS normalized radial stress, σ rr/ σ o , at r = 1.17R through the plate thickness and at various angular positions θ for a mesh density of NEL = 40

Fig. 21
figure 21

Three-dimensional ANSYS normalized shear stress, τ / σ o, at r = 1.17R through the plate thickness and at various angular positions θ for a mesh density of NEL = 40

Rights and permissions

Reprints and permissions

About this article

Cite this article

Lin, SJ., Quinn, S., Matthys, D.R. et al. Thermoelastic Determination of Individual Stresses in Vicinity of a Near-Edge Hole Beneath a Concentrated Load. Exp Mech 51, 797–814 (2011). https://doi.org/10.1007/s11340-010-9379-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11340-010-9379-6

Keywords

Navigation