Skip to main content
Log in

Force sensing using 3D displacement measurements in linear elastic bodies

  • Original Paper
  • Published:
Computational Mechanics Aims and scope Submit manuscript

Abstract

In cell traction microscopy, the mechanical forces exerted by a cell on its environment is usually determined from experimentally measured displacement by solving an inverse problem in elasticity. In this paper, an innovative numerical method is proposed which finds the “optimal” traction to the inverse problem. When sufficient regularization is applied, we demonstrate that the proposed method significantly improves the widely used approach using Green’s functions. Motivated by real cell experiments, the equilibrium condition of a slowly migrating cell is imposed as a set of equality constraints on the unknown traction. Our validation benchmarks demonstrate that the numeric solution to the constrained inverse problem well recovers the actual traction when the optimal regularization parameter is used. The proposed method can thus be applied to study general force sensing problems, which utilize displacement measurements to sense inaccessible forces in linear elastic bodies with a priori constraints.

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

Similar content being viewed by others

References

  1. Hall MS et al (2013) Toward single cell traction microscopy within 3D collagen matrices. Exp Cell Res 319(16):2396–2408

    Article  Google Scholar 

  2. DuFort CC, Paszek MJ, Weaver VM (2011) Balancing forces: architectural control of mechanotransduction. Nat Rev Mol Cell Biol 12(5):308–319

    Article  Google Scholar 

  3. Paszek MJ et al (2005) Tensional homeostasis and the malignant phenotype. Cancer Cell 8(3):241–254

    Article  MathSciNet  Google Scholar 

  4. Kostic A, Lynch CD, Sheetz MP (2009) Differential matrix rigidity response in breast cancer cell lines correlates with the tissue tropism. PLoS One 4(7):e6361

    Article  Google Scholar 

  5. Culver JC, Dickinson ME (2010) The effects of hemodynamic force on embryonic development. Microcirculation 17(3):164–178

    Article  Google Scholar 

  6. Tan JL et al (2003) Cells lying on a bed of microneedles: an approach to isolate mechanical force. Proc Natl Acad Sci 100(4):1484–1489

    Article  Google Scholar 

  7. Dembo M, Wang Y-L (1999) Stresses at the cell-to-substrate interface during locomotion of fibroblasts. Biophys J 76(4):2307–2316

    Article  Google Scholar 

  8. Legant WR et al (2010) Measurement of mechanical tractions exerted by cells in three-dimensional matrices. Nat Methods 7(12):969–971

    Article  Google Scholar 

  9. Griffith LG, Swartz MA (2006) Capturing complex 3D tissue physiology in vitro. Nat Rev Mol Cell Biol 7(3):211–224

    Article  Google Scholar 

  10. Cukierman E et al (2001) Taking cell-matrix adhesions to the third dimension. Science 294(5547):1708–1712

    Article  Google Scholar 

  11. Maniatty A, Zabaras N, Stelson K (1989) Finite element analysis of some inverse elasticity problems. J Eng Mech 115(6):1303–1317

    Article  Google Scholar 

  12. Zabaras N, Morellas V, Schnur D (1989) Spatially regularized solution of inverse elasticity problems using the BEM. Commun Appl Numer Methods 5(8):547–553

    Article  MATH  Google Scholar 

  13. Gjorevski N, Nelson CM (2012) Mapping of mechanical strains and stresses around quiescent engineered three-dimensional epithelial tissues. Biophys J 103(1):152–162

    Article  Google Scholar 

  14. Koch TM et al (2012) 3D traction forces in cancer cell invasion. PLoS One 7(3):e33476

    Article  Google Scholar 

  15. Butler JP et al (2002) Traction fields, moments, and strain energy that cells exert on their surroundings. Am J Physiol-Cell Physiol 282(3):C595–C605

    Article  Google Scholar 

  16. Storm C et al (2005) Nonlinear elasticity in biological gels. Nature 435(7039):191–194

    Article  Google Scholar 

  17. Franck C et al (2007) Three-dimensional full-field measurements of large deformations in soft materials using confocal microscopy and digital volume correlation. Exp Mech 47(3):427–438

    Article  Google Scholar 

  18. Hadamard J (1902) Sur les problèmes aux dérivées partielles et leur signification physique. Princet Univ Bull 13(49–52):28

    Google Scholar 

  19. Hofmann B, Scherzer O (1998) Local ill-posedness and source conditions of operator equations in Hilbert spaces. Inverse Probl 14(5):1189

    Article  MathSciNet  MATH  Google Scholar 

  20. Eremeyev VA, Lebedev LP (2013) Existence of weak solutions in elasticity. Math Mech Solids 18(2):204–217

    Article  MathSciNet  Google Scholar 

  21. Gurtin ME (1973) The linear theory of elasticity, in linear theories of elasticity and thermoelasticity. Springer, Berlin p 1-295

    Google Scholar 

  22. Desoer C, Whalen B (1963) A note on pseudoinverses. J Soc Ind Appl Math 11(2):442–447

    Article  MathSciNet  MATH  Google Scholar 

  23. Frommer A, Maass P (1999) Fast CG-based methods for Tikhonov–Phillips regularization. SIAM J Sci Comput 20(5):1831–1850

    Article  MathSciNet  MATH  Google Scholar 

  24. Ben-Israel A (2003) Generalized inverses : theory and applications. In: Greville TNE (ed), Springer-Verlag, New York

  25. Nashed MZ (1987) A new approach to classification and regularization of ill-posed operator equations. In: Engl HW, Groetsch CW (eds) Inverse and Ill-posed problems, notes and reports in mathematics in science and engineering, vol 4. Academic, New York, pp 53–75

    Google Scholar 

  26. Cheney E, Kincaid, D (2007) Numerical Mathematics and Computing. Cengage Learning

  27. Gould NI, Hribar ME, Nocedal J (2001) On the solution of equality constrained quadratic programming problems arising in optimization. SIAM J Sci Comput 23(4):1376–1395

    Article  MathSciNet  MATH  Google Scholar 

  28. Calvetti D et al (2000) Tikhonov regularization and the L-curve for large discrete ill-posed problems. J Comput Appl Math 123(1):423–446

    Article  MathSciNet  MATH  Google Scholar 

  29. Hansen PC (1992) Analysis of discrete ill-posed problems by means of the L-curve. SIAM Rev 34(4):561–580

    Article  MathSciNet  MATH  Google Scholar 

  30. Hansen PC, O’Leary DP (1993) The use of the L-curve in the regularization of discrete ill-posed problems. SIAM J Sci Comput 14(6):1487–1503

    Article  MathSciNet  MATH  Google Scholar 

  31. Hestenes MR, Stiefel E (1952) Methods of conjugate gradients for solving linear systems. NBS 49:409–436

    MathSciNet  MATH  Google Scholar 

  32. Bangerth W et al (2015) The deal.II Library, version 8.2. Arch Numer Softw 3:1–8. doi:10.11588/ans.2015.100.18031

  33. Tomei AA et al (2009) 3D collagen cultures under well-defined dynamic strain: a novel strain device with a porous elastomeric support. Biotechnol Bioeng 103(1):217–225

    Article  MathSciNet  Google Scholar 

  34. Barocas VH, Moon AG, Tranquillo RT (1995) The fibroblast-populated collagen microsphere assay of cell traction force—part 2: measurement of the cell traction parameter. J Biomech Eng 117(2):161–170

    Article  Google Scholar 

  35. Wakitani S et al (1998) Repair of large full-thickness articular cartilage defects with allograft articular chondrocytes embedded in a collagen gel. Tissue Eng 4(4):429–444

    Article  Google Scholar 

  36. Eshelby JD (1957) The determination of the elastic field of an ellipsoidal inclusion, and related problems. In: Proceedings of the Royal Society of London A: mathematical, physical and engineering sciences. The Royal Society

  37. Mura T (2012) Micromechanics of defects in solids, vol 3. Springer Science & Business Media, Berlin

    Google Scholar 

  38. Trickey WR et al (2006) Determination of the Poisson’s ratio of the cell: recovery properties of chondrocytes after release from complete micropipette aspiration. J Biomech 39(1):78–87

    Article  MathSciNet  Google Scholar 

  39. Lindström SB et al (2013) Finite-strain, finite-size mechanics of rigidly cross-linked biopolymer networks. Soft Matter 9(30):7302–7313

    Article  Google Scholar 

  40. Staunton JR (2014) Mechanics of cancer cells in 3D microenvironments. Arizona State University, Tempe

    Google Scholar 

  41. Meng C, Heltsley W, Pollard DD (2012) Evaluation of the Eshelby solution for the ellipsoidal inclusion and heterogeneity. Comput Geosci 40:40–48

  42. Geuzaine C, Remacle JF (2009) Gmsh: a 3-D finite element mesh generator with built-in pre-and post-processing facilities. Int J Numer Meth Eng 79(11):1309–1331

    Article  MathSciNet  MATH  Google Scholar 

  43. Vader D et al (2009) Strain-induced alignment in collagen gels. PloS one 4(6):e5902

    Article  Google Scholar 

  44. Janmey PA et al (2007) Negative normal stress in semiflexible biopolymer gels. Nat Mater 6(1):48–51

    Article  Google Scholar 

  45. Mofrad MR (2009) Rheology of the cytoskeleton. Ann Rev Fluid Mech 41:433–453

    Article  MATH  Google Scholar 

  46. Karcher H et al (2003) A three-dimensional viscoelastic model for cell deformation with experimental verification. Biophys J 85(5):3336–3349

    Article  Google Scholar 

  47. Yoshigi M et al (2005) Mechanical force mobilizes zyxin from focal adhesions to actin filaments and regulates cytoskeletal reinforcement. J Cell Biol 171(2):209–215

    Article  Google Scholar 

  48. De Borst R et al (2012) Nonlinear finite element analysis of solids and structures. Wiley, New York

    Book  MATH  Google Scholar 

  49. Bertsekas DP (1999) Nonlinear programming. Athena Scientific, Belmont

    MATH  Google Scholar 

  50. Jacoby SL, Kowalik JS, Pizzo JT (1972) Iterative methods for nonlinear optimization problems. Prentice-Hall, Englewood Cliffs

    MATH  Google Scholar 

  51. Hager WW, Zhang H (2006) A survey of nonlinear conjugate gradient methods. Pac J Optim 2(1):35–58

    MathSciNet  MATH  Google Scholar 

  52. Boas RP, Boas HP (1996) A primer of real functions, vol 13. Cambridge University Press, Cambridge

    MATH  Google Scholar 

Download references

Acknowledgments

XZF is partially supported by the National Center for Research Resources (5R21RR025801-03). Both authors give thanks to Dr. Mingming Wu and Matthew S. Hall for motivation of the inverse problem in cell traction microscopy. XZF gives thanks to Dr. Timothy J. Healey and Chenxi Wu for helpful discussions on the proof of the convergence theorem in the appendix.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xinzeng Feng.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (docx 1592 KB)

Appendix

Appendix

1.1 General cases

In this part, we extend our discussion to general boundary value problem in linear elasticity:

$$\begin{aligned} \nabla \cdot {\varvec{\sigma }} +{\varvec{f}}=0,\quad \hbox { in }\Omega ,\quad {\varvec{u}}|_{\Gamma _{u} } =\bar{{\varvec{u}}},\quad {\varvec{\sigma }}\cdot {\varvec{n}}|_{\Gamma _{t} } =\bar{{\varvec{t}}} \end{aligned}$$
(45)

where \({\varvec{f}}\) is the body force density, \(\bar{{\varvec{u}}}\) is the prescribed displacements on \(\Gamma _{u} \) and \(\bar{{\varvec{t}}}\) is the unknown traction on \(\Gamma _{t} \).

Denote the solution to (45) as \({\varvec{u}}\left[ {\bar{{\varvec{t}}}} \right] \). According to the principle of superposition,

$$\begin{aligned} {\varvec{u}}\left[ {\bar{{\varvec{t}}}} \right] ={\varvec{u}}_{0} +{\varvec{u}}^{\circ }\left[ {\bar{{\varvec{t}}}} \right] \end{aligned}$$
(46)

where \({\varvec{u}}_{0} \) is the solution to problem

$$\begin{aligned} \nabla \cdot {\varvec{\sigma }} +{\varvec{f}}=0,\quad \hbox { in }\Omega ,\quad {\varvec{u}}|_{\Gamma _{u} } =\bar{{\varvec{u}}},\quad {\varvec{\sigma }}\cdot {\varvec{n}}|_{\Gamma _{t} } =0 \end{aligned}$$
(47)

and \({\varvec{u}}^{\circ }\left[ {\bar{{\varvec{t}}}} \right] \) is the solution to problem

$$\begin{aligned} \nabla \cdot {\varvec{\sigma }}=0,\quad \hbox { in }\Omega ,\quad {\varvec{u}}|_{\Gamma _{u} } =0,\quad {\varvec{\sigma }}\cdot {\varvec{n}}|_{\Gamma _{t} } =\bar{{\varvec{t}}}. \end{aligned}$$
(48)

The maps of prediction to (45) and (48) are respectively \(T\left( {\bar{{\varvec{t}}}} \right) =T_{m} \left( {{\varvec{u}}\left[ {\bar{{\varvec{t}}}} \right] } \right) \) and \(T^{\circ }\left( {\bar{{\varvec{t}}}} \right) =T_{m} \left( {{\varvec{u}}^{\circ }\left[ {\bar{{\varvec{t}}}} \right] } \right) \), where \(T_{m} \) is the map of sampling defined by

$$\begin{aligned} T_{m} \left( {{\varvec{w}}} \right) \equiv \left[ {{\begin{array}{l} {{\varvec{w}}({{\varvec{p}}}_{1} )} \\ \vdots \\ {{\varvec{w}}({{\varvec{p}}}_{M} )} \\ \end{array} }} \right] ,\quad \forall {\varvec{w}}\in {{\varvec{C}}}\left( \Omega \right) . \end{aligned}$$
(49)

Therefore, (46) leads to

$$\begin{aligned} T\left( {\bar{{\varvec{t}}}} \right) =T_{m} \left( {{\varvec{u}}_{0} } \right) +T^{o}\left( {\bar{{\varvec{t}}}} \right) \end{aligned}$$
(50)

with which the original inverse problem \(T\left( {\bar{{\varvec{t}}}} \right) =\hat{{\varvec{U}}}_{m} \) is found to be equivalent to

$$\begin{aligned} T^{\circ }\left( {\bar{{\varvec{t}}}} \right) =\hat{{\varvec{U}}}_{m} -T_{m} \left( {{\varvec{u}}_{0} } \right) . \end{aligned}$$
(51)

1.2 Proof of Theorem in Sect. 2

Consider the boundary value problem (48) with homogeneous displacement boundary and vanishing body force. Denote the measurement points in the interior of \(\Omega \) as \({\varvec{p}}_{i} ,i=1,\ldots , M\) and the corresponding map of prediction as T. Since T is a continuous linear transformation between Hilbert spaces \({\varvec{W}}_{f} \), i.e. \({\varvec{L}}\left( {\Gamma _{t} } \right) \), and \(R^{3M}\), there exists a unique adjoint transformation \(T^{*}:R^{3M}\rightarrow {\varvec{W}}_{f} \) which is also linear and continuous. By definition, \(\forall {\varvec{v}}_{m} \in R^{3M},\bar{{\varvec{t}}}\in {{\varvec{W}}}_{f} \), \(T^{*}\) satisfies

$$\begin{aligned} T\left( {\bar{{\varvec{t}}}} \right) \cdot {\varvec{v}}_{m} =\int \limits _{\Gamma _{t} } {\bar{{\varvec{t}}}\cdot T^{*}\left( {{\varvec{v}}_{m} } \right) dA}. \end{aligned}$$
(52)

On the other hand, the solution to (48) is given by

$$\begin{aligned} {\varvec{u}}[\bar{{\varvec{t}}}]\left( {{\varvec{x}}} \right) =\int \limits _{\Gamma _{t} } {{\varvec{G}}\left( {{\varvec{x}},{{\varvec{y}}}} \right) \bar{{\varvec{t}}}\left( {{\varvec{y}}} \right) dA} \end{aligned}$$
(53)

where the \(i{-}j\) component of matrix \({\varvec{G}}\left( {{\varvec{x}},{{\varvec{y}}}} \right) \) gives the displacement at point \({\varvec{x}}\) in the \(i^{{\textit{th}}}\) direction caused by a unit force in the \(j^{{\textit{th}}}\) direction at point \({\varvec{y}}\) on \(\Gamma _{t} \), which we denote as \(G_{ij} \left( {{\varvec{x}},{{\varvec{y}}}} \right) \).

Notice \(T\left( {\bar{{\varvec{t}}}} \right) =T_{m} \left( {{\varvec{u}}\left[ {\bar{{\varvec{t}}}} \right] } \right) \), (53) leads to

$$\begin{aligned} T\left( {\bar{{\varvec{t}}}} \right) \cdot {\varvec{v}}_{m}= & {} \int \limits _{\Gamma _{t} } {T_{m} \left( {{\varvec{G}}\left( {\cdot , {\varvec{y}}} \right) \bar{{\varvec{t}}}\left( {{\varvec{y}}} \right) } \right) \cdot {\varvec{v}}_{m} dA} \nonumber \\= & {} \int \limits _{\Gamma _{t} } {\bar{{\varvec{t}}}\left( {{\varvec{y}}} \right) \cdot \left( {\sum _{i=1}^M {{\varvec{G}}\left( {{\varvec{p}}_{i}, {\varvec{y}}} \right) \cdot {\varvec{u}}_{i} } } \right) dA.} \end{aligned}$$
(54)

where \({\varvec{v}}_{m} =\left( {{\varvec{u}}_{1}^{T},\ldots ,{\varvec{u}}_{M}^{T} } \right) \in R^{3M}\).

Due to the uniqueness of the adjoint operator, (52) and (54) gives

$$\begin{aligned} T^{*}\left( {{\varvec{v}}_{m} ;{\varvec{y}}} \right) =\sum _{i=1}^M {{\varvec{G}}\left( {{\varvec{p}}_{i}, {\varvec{y}}} \right) \cdot {\varvec{u}}_{i} },\quad \forall {\varvec{y}}\in \Gamma _{t}. \end{aligned}$$
(55)

Here \(T^{*}\left( {{\varvec{v}}_{m} ;{\varvec{y}}} \right) \) denotes the evaluation of \(T^{*}\left( {{\varvec{v}}_{m} } \right) \) at point \({\varvec{y}}\in \Gamma _{t} \). The range of \(T^{*}\) is therefore given by

$$\begin{aligned} {\mathcal {R}}\left( {T}^{*} \right) =span\left\{ {\bar{{\varvec{t}}}_{ij}, j=1,2,3} \right\} _{i=1}^{M} \end{aligned}$$
(56)

where

$$\begin{aligned} \bar{{\varvec{t}}}_{ij} \left( {{\varvec{y}}} \right) ={\varvec{G}}\left( {{\varvec{p}}_{i}, {\varvec{y}}} \right) \cdot {\varvec{e}}_{j}. \end{aligned}$$
(57)

According to [22], the error function vanishes for any traction in \({\mathcal {R}}\left( {T}^{*} \right) \), i.e.

$$\begin{aligned} E\left( {\bar{{\varvec{t}}}} \right) \equiv \bar{{\varvec{t}}}-T^{\dag }\left( {T\left( {\bar{{\varvec{t}}}} \right) } \right) =0,\quad \forall \bar{{\varvec{t}}}\in {\mathcal {R}}\left( {T}^{*} \right) \end{aligned}$$
(58)

More generally, knowing that \({\mathcal {R}}\left( {T}^{*} \right) \) is closed in \({\varvec{W}}_{f} \), for any \(\bar{{\varvec{t}}}\in {{\varvec{W}}}_{f} \), there exists a unique decomposition \(\bar{{\varvec{t}}}=\bar{{\varvec{t}}}_{1} +\bar{{\varvec{t}}}_{2} \) where \(\bar{{\varvec{t}}}_{1} \in {\mathcal {R}}\left( {T}^{*} \right) \) and \(\bar{{\varvec{t}}}_{2} \in {\mathcal {R}}\left( {T}^{*} \right) ^{\perp }=Ker\left( T \right) \). Therefore, \(T\left( {\bar{{\varvec{t}}}} \right) =T\left( {\bar{{\varvec{t}}}_{1} } \right) \) and

$$\begin{aligned} E\left( {\bar{{\varvec{t}}}} \right) =\bar{{\varvec{t}}}-T\left( {T\left( {\bar{{\varvec{t}}}_{1} } \right) } \right) =\bar{{\varvec{t}}}_{2}. \end{aligned}$$
(59)

Since \(\left\| {\bar{{\varvec{t}}}} \right\| _{L^{2}}^{2} =\left\| {\bar{{\varvec{t}}}_{1} } \right\| _{L^{2}}^{2} +\left\| {\bar{{\varvec{t}}}_{2} } \right\| _{L^{2}}^{2} \), it is known from (59) that E is a bounded linear transformation with \(\left\| E \right\| \le 1\).

Now consider an infinite series of measurement points \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \). We call \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) to be locally dense in \(\Omega \) if \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) is dense in a sub-domain \(\Omega _{0} \subset \Omega \) with positive measure, i.e. the closure \(\overline{\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } } \) of \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) contains \(\Omega _{0} \) and \(meas\left( {\Omega _{0} } \right) >0\).

For any given \(M>0\), we denote the map of prediction of the inverse problem given partial measurement points \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{M} \) as \(T_{M} \) and the corresponding error function as \(E_{M} \left( {\bar{{\varvec{t}}}} \right) \equiv \left( {I-T_{M}^{\dagger } \circ T_{M} } \right) \left( {\bar{{\varvec{t}}}} \right) \). We claim that if \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) is locally dense in \(\Omega \), then

$$\begin{aligned} \left\| {E_{M} \left( {\bar{{\varvec{t}}}} \right) } \right\| _{L^{2}} \rightarrow 0,\quad \hbox {as }M\rightarrow \infty . \end{aligned}$$
(60)

In order to prove (60), we need the following lemmas:

Lemma 1

If a real analytic function \({\varvec{u}}\) on \(\Omega \) vanishes on a set of positive measure, then \({\varvec{u}}\) is identically zero [52].

Lemma 2

If \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) is locally dense in \(\Omega \), then \({\mathcal {R}}_{\infty }^{\bot } =\left\{ {\mathbf {0}} \right\} \). Here \({\mathcal {R}}_{\infty } \equiv span\left\{ {\bar{{\varvec{t}}}_{ij}, j=1,2,3} \right\} _{i=1}^{\infty } \) where \(\bar{{\varvec{t}}}_{ij} \) is defined in (57).

Proof

Consider an arbitrary element \(\bar{{\varvec{t}}}^{\prime }\in {\mathcal {R}}_{\infty }^{\bot } \) which by definition satisfies

$$\begin{aligned} \int \limits _{\Gamma _{t} } {\bar{{\varvec{t}}}^{\prime }\cdot \bar{{\varvec{t}}}_{ij} dA} =0, \quad \forall i\ge 1,\quad j=1,2,3 \end{aligned}$$
(61)

Notice

$$\begin{aligned} \int \limits _{\Gamma _{t} } {\bar{{\varvec{t}}}^{\prime }\cdot \bar{{\varvec{t}}}_{ij} dA}= & {} \int \limits _{\Gamma _{t} } {\bar{{\varvec{t}}}^{\prime }\left( {{\varvec{y}}} \right) \cdot {\varvec{G}}\left( {{\varvec{p}}_{i}, {\varvec{y}}} \right) \cdot {\varvec{e}}_{j} dA} \end{aligned}$$
(62)
$$\begin{aligned}= & {} {\varvec{e}}_{j} \cdot \int \limits _{\Gamma _{t} } {{\varvec{G}}\left( {{\varvec{p}}_{i}, {\varvec{y}}} \right) \bar{{\varvec{t}}}^{\prime }\left( {{\varvec{y}}} \right) dA} ={\varvec{e}}_{j} \cdot {\varvec{u}}\left[ \bar{{\varvec{t}}}^{\prime } \right] ({\varvec{p}}_{i} ).\nonumber \\ \end{aligned}$$
(63)

(61) indicates that the elastic displacement \({\varvec{u}}\left[ \bar{{\varvec{t}}}^{\prime } \right] \) to (48) vanishes at \({\varvec{p}}_{i} \) for \(i\ge 1\). On the other hand, according to Gurtin [21], \({\varvec{u}}\left[ \bar{{\varvec{t}}}^{\prime } \right] \) is a real analytic function in \(\Omega \). Due to lemma 1, therefore, the elastic solution vanishes everywhere in \(\Omega \), i.e. \({\varvec{u}}\left[ \bar{{\varvec{t}}}^{\prime } \right] =\mathbf {0}\). Further, notice that the material is assumed to be compressible, we have \({\varvec{\sigma }}\left[ \bar{{\varvec{t}}}^{\prime } \right] =\mathbf {0}\) and hence \(\bar{{\varvec{t}}}^{\prime }={\varvec{\sigma }}\left[ \bar{{\varvec{t}}}^{\prime } \right] \cdot {\varvec{n}}=\mathbf {0}\) on \(\Gamma _{t} \).

Corollary

\(\overline{{\mathcal {R}}_{\infty } } =\left\{ \mathbf{0} \right\} ^{\bot }={\varvec{W}}_{f} \), i.e., \(\left\{ {\bar{{\varvec{t}}}_{ij} ,j=1,2,3,i\ge 1} \right\} \) is dense in \({\varvec{W}}_{f} \).

We are now in a position to prove (60), which is summarized in the following theorem:

Theorem

If \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) is locally dense in \(\Omega \), then \(\forall \bar{{\varvec{t}}}\in {{\varvec{W}}}_{f} \), \(\left\| {E_{M} \left( {\bar{{\varvec{t}}}} \right) } \right\| _{L^{2}} \rightarrow 0\) as \(M\rightarrow \infty \).

Proof

According to the above corollary, \(\left\{ \bar{{\varvec{t}}}_{ij}, j=1,2,3,\right. \left. i\ge 1 \right\} \) is dense in \({\varvec{W}}_{f} \) given that \(\left\{ {{\varvec{p}}_{i} } \right\} _{i=1}^{\infty } \) is locally dense in \(\Omega \). As a result, for any \(\bar{{\varvec{t}}}\in {{\varvec{W}}}_{f} \) and \(\varepsilon >0\), there exists \(\left\{ {{\varvec{u}}_{i} } \right\} _{i=1}^{\infty } \) and \(M_{\varepsilon } \) such that \(\left\| {\bar{{\varvec{t}}}-\bar{{\varvec{t}}}_{p}^{M} } \right\| _{L^{2}} <\varepsilon \) for \(\forall M>M_{\varepsilon } \) where \(\bar{{\varvec{t}}}_{p}^{M} =\sum \limits _{i=1}^M {{\varvec{G}}\left( {{\varvec{p}}_{i}, {\varvec{y}}} \right) \cdot {\varvec{u}}_{i} } \). Note \(\bar{{\varvec{t}}}_{p}^{M} \in {\mathcal {R}}\left( {T_{M}^{{*}} } \right) \), we have \(E_{M} \left( {\bar{{\varvec{t}}}_{p}^{M} } \right) =0\) and further, due to linearity of the error function,

$$\begin{aligned} E_{M} \left( {\bar{{\varvec{t}}}} \right) =E_{M} \left( {\bar{{\varvec{t}}}-\bar{{\varvec{t}}}_{p}^{M} } \right) . \end{aligned}$$
(64)

From our earlier discussion, \(\left\| {E_{M} } \right\| \le 1\) is uniformly bounded. Therefore, (64) leads to that \(\left\| {E_{M} \left( {\bar{{\varvec{t}}}} \right) } \right\| _{L^{2}} \le \left\| {\bar{{\varvec{t}}}-\bar{{\varvec{t}}}_{p}^{M} } \right\| _{L^{2}} <\varepsilon \), which is true for any \(M>M_{\varepsilon } \).

To this end, we’ve proved that \(\forall \bar{{\varvec{t}}}\in {{\varvec{W}}}_{f} \), \(\left\| {E_{M} \left( {\bar{{\varvec{t}}}} \right) } \right\| _{L^{2}} \rightarrow 0\), as \(M\rightarrow \infty \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Feng, X., Hui, CY. Force sensing using 3D displacement measurements in linear elastic bodies. Comput Mech 58, 91–105 (2016). https://doi.org/10.1007/s00466-016-1283-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-016-1283-1

Keywords

Navigation