Skip to main content

Transport Properties

  • Chapter
  • First Online:
Helical Wormlike Chains in Polymer Solutions

Abstract

This chapter deals with the classical hydrodynamic theory of steady-state transport properties, such as the translational friction and diffusion coefficients and intrinsic viscosity, of the unperturbed HW chain, including the KP wormlike chain as a special case, on the basis of the cylinder and touched-bead models. An analysis of experimental data is made from various points of view, which are based on the present theory, especially for flexible polymers. In the same spirit as that in Chap. 5, use is then made of experimental data obtained for several flexible polymers in the \(\varTheta\) state over a wide range of molecular weight, including the oligomer region, and also for typical semiflexible polymers. As a result, it is pointed out that there still remain several unsolved problems for flexible polymers even in the unperturbed state. It is convenient to begin by giving a general consideration of some aspects of polymer hydrodynamics which leads to the adoption of the present hydrodynamic models.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. H. Yamakawa, Modern Theory of Polymer Solutions (Harper & Row, New York, 1971). Its electronic edition is available on-line at the URL: http://hdl.handle.net/2433/50527

  2. M. Doi, S.F. Edwards, The Theory of Polymer Dynamics (Clarendon Press, Oxford, 1986)

    Google Scholar 

  3. J.G. Kirkwood, J. Riseman, J. Chem. Phys. 16, 565 (1948)

    Article  CAS  Google Scholar 

  4. J.G. Kirkwood, J. Polym. Sci. 12, 1 (1954)

    Article  CAS  Google Scholar 

  5. R.E. DeWames, W.F. Hall, M.C. Shen, J. Chem. Phys. 46, 2782 (1967)

    Article  CAS  Google Scholar 

  6. R. Zwanzig, J. Kiefer, G.H. Weiss, Proc. Natl. Acad. Sci. U. S. A. 60, 381 (1968)

    Article  CAS  Google Scholar 

  7. H. Yamakawa, Ann. Rev. Phys. Chem. 25, 179 (1974)

    Article  Google Scholar 

  8. J. Rotne, S. Prager, J. Chem. Phys. 50, 4831 (1969)

    Article  CAS  Google Scholar 

  9. H. Yamakawa, J. Chem. Phys. 53, 436 (1970)

    Article  CAS  Google Scholar 

  10. T. Yoshizaki, H. Yamakawa, J. Chem. Phys. 73, 578 (1980)

    Article  CAS  Google Scholar 

  11. H. Yamakawa, G. Tanaka, J. Chem. Phys. 57, 1537 (1972)

    Article  CAS  Google Scholar 

  12. J. García de la Torre, V.A. Bloomfield, Q. Rev. Biophys. 14, 81 (1981)

    Google Scholar 

  13. H. Yamakawa, Ann. Rev. Phys. Chem. 35, 23 (1984)

    Article  CAS  Google Scholar 

  14. S.F. Edwards, G.J. Papadopoulos, J. Phys. A 1, 173 (1968)

    Article  CAS  Google Scholar 

  15. S.F. Edwards, M.A. Oliver, J. Phys. A 4, 1 (1971)

    Article  Google Scholar 

  16. J.M. Burgers, Second Report on Viscosity and Plasticity of the Amsterdam Academy of Sciences, Chap. 3 (North-Holland, Amsterdam, 1938)

    Google Scholar 

  17. C.-M. Tchen, J. Appl. Phys. 25, 463 (1954)

    Article  CAS  Google Scholar 

  18. S. Broersma, J. Chem. Phys. 32, 1632 (1960)

    Article  CAS  Google Scholar 

  19. H. Yamakawa, M. Fujii, Macromolecules 6, 407 (1973)

    Article  CAS  Google Scholar 

  20. H. Yamakawa, M. Fujii, Macromolecules 7, 128 (1974)

    Article  CAS  Google Scholar 

  21. J. Shimada, H. Yamakawa, Macromolecules 9, 583 (1976)

    Article  CAS  Google Scholar 

  22. H. Yamakawa, T. Yoshizaki, Macromolecules 12, 32 (1979)

    Article  CAS  Google Scholar 

  23. H. Yamakawa, T. Yoshizaki, Macromolecules 13, 633 (1980)

    Article  CAS  Google Scholar 

  24. H. Yamakawa, Macromolecules 16, 1928 (1983)

    Article  CAS  Google Scholar 

  25. J.E. Hearst, W.H. Stockmayer, J. Chem. Phys. 37, 1425 (1962)

    Article  CAS  Google Scholar 

  26. H. Yamakawa, J. Shimada, M. Fujii, J. Chem. Phys. 68, 2140 (1978)

    Article  CAS  Google Scholar 

  27. T. Yoshizaki, H. Yamakawa, J. Chem. Phys. 72, 57 (1980)

    Article  CAS  Google Scholar 

  28. T. Yoshizaki, I. Nitta, H. Yamakawa, Macromolecules 21, 165 (1988)

    Article  CAS  Google Scholar 

  29. H. Yamakawa, T. Yoshizaki, J. Chem. Phys. 78, 572 (1983)

    Article  CAS  Google Scholar 

  30. H. Yamakawa, J. Yamaki, J. Chem. Phys. 57, 1542 (1972)

    Article  Google Scholar 

  31. T. Yamada, T. Yoshizaki, H. Yamakawa, Macromolecules 25, 377 (1992)

    Article  CAS  Google Scholar 

  32. J.G. Kirkwood, Recl. Trav. Chim. 68, 649 (1949)

    Article  CAS  Google Scholar 

  33. H. Yamakawa, Macromolecules 8, 339 (1975)

    Article  CAS  Google Scholar 

  34. J.G. Kirkwood, P.L. Auer, J. Chem. Phys. 19, 281 (1951)

    Article  CAS  Google Scholar 

  35. R. Ullman, J. Chem. Phys. 40, 2422 (1964)

    Article  CAS  Google Scholar 

  36. P.L. Auer, C.S. Gardner, J. Chem. Phys. 23, 1546 (1955)

    Article  CAS  Google Scholar 

  37. H. Brenner, Chem. Eng. Sci. 19, 631 (1964)

    Article  CAS  Google Scholar 

  38. T. Yoshizaki, H. Yamakawa, J. Chem. Phys. 88, 1313 (1988)

    Article  CAS  Google Scholar 

  39. Y. Einaga, H. Koyama, T. Konishi, H. Yamakawa, Macromolecules 22, 3419 (1989)

    Article  CAS  Google Scholar 

  40. T. Konishi, T. Yoshizaki, T. Saito, Y. Einaga, H. Yamakawa, Macromolecules 23, 290 (1990)

    Article  CAS  Google Scholar 

  41. T. Konishi, T. Yoshizaki, H. Yamakawa, Macromolecules 24, 5614 (1991)

    Article  CAS  Google Scholar 

  42. F. Abe, Y. Einaga, H. Yamakawa, Macromolecules 26, 1891 (1993)

    Article  CAS  Google Scholar 

  43. I. Suda, Y. Tominaga, M. Osa, T. Yoshizaki, H. Yamakawa, Macromolecules 33, 9322 (2000)

    Article  CAS  Google Scholar 

  44. Y. Fujii, Y. Tamai, T. Konishi, H. Yamakawa, Macromolecules 24, 1608 (1991)

    Article  CAS  Google Scholar 

  45. F. Abe, K. Horita, Y. Einaga, H. Yamakawa, Macromolecules 27, 725 (1994)

    Article  CAS  Google Scholar 

  46. N. Sawatari, T. Konishi, T. Yoshizaki, H. Yamakawa, Macromolecules 28, 1089 (1995)

    Article  CAS  Google Scholar 

  47. M. Kamijo, F. Abe, Y. Einaga, H. Yamakawa, Macromolecules 28, 1095 (1995)

    Article  CAS  Google Scholar 

  48. F. Abe, Y. Einaga, H. Yamakawa, Macromolecules 24, 4423 (1991)

    Article  CAS  Google Scholar 

  49. T. Yamada, H. Koyama, T. Yoshizaki, Y. Einaga, H. Yamakawa, Macromolecules 26, 2566 (1993)

    Article  CAS  Google Scholar 

  50. M. Kuwata, H. Murakami, T. Norisuye, H. Fujita, Macromolecules 17, 2731 (1984)

    Article  CAS  Google Scholar 

  51. H. Murakami, T. Norisuye, H. Fujita, Macromolecules 13, 345 (1980)

    Article  CAS  Google Scholar 

  52. J.E. Godfrey, Biophys. Chem. 5, 285 (1976)

    Article  CAS  Google Scholar 

  53. J.E. Godfrey, H. Eisenberg, Biophys. Chem. 5, 301 (1976)

    Article  CAS  Google Scholar 

  54. D. Jolly, H. Eisenberg, Biopolymers 15, 61 (1976)

    Article  CAS  Google Scholar 

  55. T. Yanaki, T. Norisuye, H. Fujita, Macromolecules 13, 1462 (1980)

    Article  CAS  Google Scholar 

  56. H. Kusanagi, H. Tadokoro, Y. Chatani, Polym. J. 9, 181 (1977)

    Article  CAS  Google Scholar 

  57. K. Dehara, T. Yoshizaki, H. Yamakawa, Macromolecules 26, 5137 (1993)

    Article  CAS  Google Scholar 

  58. T. Arai, N. Sawatari, T. Yoshizaki, Y. Einaga, H. Yamakawa, Macromolecules 29, 2309 (1996)

    Article  CAS  Google Scholar 

  59. M.T. Record Jr., C.P. Woodbury, R.B. Inman, Biopolymers 14, 393 (1975)

    Article  CAS  Google Scholar 

  60. M. Osa, F. Abe, T. Yoshizaki, Y. Einaga, H. Yamakawa, Macromolecules 29, 2302 (1996)

    Article  CAS  Google Scholar 

  61. T. Norisuye, Prog. Polym. Sci. 18, 543 (1993)

    Article  CAS  Google Scholar 

  62. Y. Tamai, T. Konishi, Y. Einaga, M. Fujii, H. Yamakawa, Macromolecules 23, 4067 (1990)

    Article  CAS  Google Scholar 

  63. B.H. Zimm, J. Chem. Phys. 24, 269 (1956)

    Article  CAS  Google Scholar 

  64. M. Fixman, J. Chem. Phys. 42, 3831 (1965)

    Article  CAS  Google Scholar 

  65. C.W. Pyun, M. Fixman, J. Chem. Phys. 42, 3838 (1965)

    Article  CAS  Google Scholar 

  66. B.H. Zimm, Macromolecules 13, 592 (1980)

    Article  CAS  Google Scholar 

  67. M. Fixman, J. Chem. Phys. 80, 6324 (1984); 84, 4085 (1986)

    Article  CAS  Google Scholar 

  68. M. Fixman, Faraday Discuss. 83, 199 (1987); J. Chem. Phys. 89, 2442 (1988)

    Article  CAS  Google Scholar 

  69. H. Yamakawa, T. Yoshizaki, J. Chem. Phys. 91, 7900 (1989)

    Article  CAS  Google Scholar 

  70. P. Rempp, J. Polym. Sci. 23, 83 (1957)

    Article  CAS  Google Scholar 

  71. E.D. von Meerwall, S. Amelar, M.A. Smeltzly, T.P. Lodge, Macromolecules 22, 295 (1989)

    Article  Google Scholar 

  72. A. E. Dunstan, J. Chem. Soc. Lond. 85, 817 (1904)

    Article  CAS  Google Scholar 

  73. von W. Herz, W. Rathmann, Z. Elektrochem. 19, 589 (1913)

    Google Scholar 

  74. M. Fixman, J. Chem. Phys. 92, 6858 (1990)

    Article  CAS  Google Scholar 

  75. M. Fujii, H. Yamakawa, Macromolecules 8, 792 (1975)

    Article  CAS  Google Scholar 

  76. V.A. Bloomfield, B.H. Zimm, J. Chem. Phys. 44, 315 (1966)

    Article  CAS  Google Scholar 

  77. M. Fukatsu, M. Kurata, J. Chem. Phys. 44, 4539 (1966)

    Article  CAS  Google Scholar 

  78. R. Zwanzig, J. Chem. Phys. 45, 1858 (1966)

    Article  CAS  Google Scholar 

  79. E. Paul, R.M. Mazo, J. Chem. Phys. 48, 2378 (1968); 51, 1102 (1969)

    Article  CAS  Google Scholar 

  80. G. Tanaka, H. Yamakawa, Polym. J. 4, 446 (1973)

    Article  CAS  Google Scholar 

  81. L.V. Crawford, P.H. Bleck, Virology 24, 388 (1964)

    Article  CAS  Google Scholar 

  82. J. Vinograd, J. Lebowitz, R. Radloff, R. Watson, P. Laipis, Proc. Natl. Acad. Sci. U. S. A. 53, 1104 (1965)

    Article  CAS  Google Scholar 

  83. L.V. Crawford, J. Mol. Biol. 13, 362 (1965)

    Article  CAS  Google Scholar 

  84. C. Bode, A.D. Kaiser, J. Mol. Biol. 14, 399 (1965)

    Article  CAS  Google Scholar 

  85. I.B. David, D.R. Wolstenholme, J. Mol. Biol. 28, 233 (1967)

    Article  Google Scholar 

  86. A. Opschoor, P.H. Pouwels, C.M. Knijnenburg, J.B.T. Aten, J. Mol. Biol. 37, 13 (1968)

    Article  CAS  Google Scholar 

  87. J.R. Dawson, J.A. Harpst, Biopolymers 10, 2499 (1971)

    Article  CAS  Google Scholar 

  88. D.A. Ostrander, H.B. Gray Jr., Biopolymers 12, 1387 (1973)

    Article  CAS  Google Scholar 

  89. H. Yamakawa, T. Yoshizaki, M. Fujii, Macromolecules 10, 934 (1977)

    Article  CAS  Google Scholar 

  90. T. Norisuye, M. Motowoka, H. Fujita, Macromolecules 12, 320 (1979)

    Article  CAS  Google Scholar 

  91. G.B. Jeffery, Proc. R. Soc. Lond. Ser. A 102, 161 (1922)

    Article  Google Scholar 

  92. F. Perrin, J. Phys. Rad. 7, 1 (1936)

    Article  CAS  Google Scholar 

  93. S. Broersma, J. Chem. Phys. 74, 6989 (1981)

    Article  Google Scholar 

  94. R. Simha, J. Phys. Chem. 44, 25 (1940)

    Article  CAS  Google Scholar 

  95. N. Saito, J. Phys. Soc. Jpn. 6, 297 (1951)

    Article  CAS  Google Scholar 

  96. A. Einstein, Ann. Phys. (Leipzig) 19, 289 (1906); 34, 591 (1911)

    Article  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 1: Transport Coefficients of Spheroid-Cylinders

In this appendix we evaluate the translational and rotatory friction (or diffusion) coefficients and intrinsic viscosity of the spheroid-cylinder defined in Sect. 6.2.1 and depicted in Fig. 6.2 [27]. We introduce external (e x , e y , e z ) and molecular (e 1, e 2, e 3) Cartesian coordinate systems, choosing the center of mass of the body as the origin of the latter. The superscript (e) is used to indicate vectors and tensors expressed in the external system, and no superscript is used for those in the molecular system. The spatial configuration of the body may be determined by the vector position \(\mathbf{R}_{\mathrm{c}}^{\mathrm{(e)}} \equiv \mathbf{R}_{\mathrm{c}} = (x_{\mathrm{c}}\), y c, z c) of the center of mass in the external system and the Euler angles \(\varTheta = (\alpha\), β, γ) defining the orientation of the molecular system with respect to the external system. The matrix transforming the external coordinates to the molecular coordinates, which we denote by A, is then identical to the matrix Q given by Eq. (4.96) with (α, β, γ) in place of (\(\tilde{\theta }\), \(\tilde{\phi }\), \(\tilde{\psi }\)).

Now let U c be the instantaneous velocity of the center of mass of the body, let \(\boldsymbol{\Omega }\) be its instantaneous angular velocity, and let v 0 be the unperturbed flow field of a solvent. Under the nonslip boundary condition, the frictional force f(r) exerted by the unit area at r of the surface of the body satisfies the integral equation

$$\displaystyle{ 8\pi \eta _{0}{\bigl [\mathbf{U}_{\mathrm{c}} +\boldsymbol{ \Omega } \times \mathbf{r}_{1} -\mathbf{v}^{0}(\mathbf{r}_{ 1})\bigr ]} =\int _{\mathrm{S}}\mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{2}) \cdot \mathbf{f}(\mathbf{r}_{2})d\mathbf{r}_{2}\,, }$$
(6.154)

where r 1 and r 2 are the vector positions of two arbitrary points on the surface of the body, \(\mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{2})\) is defined by

$$\displaystyle{ \mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{2}) = 8\pi \eta _{0}\mathbf{T}(\mathbf{r}_{1} -\mathbf{r}_{2}) }$$
(6.155)

as in Eq. (6.58), and the integration in Eq. (6.154) is carried out over the surface of the body. If we define the inverse \(\mathbf{K}^{-1}(\mathbf{r}_{1},\mathbf{r}_{2})\) by

$$\displaystyle\begin{array}{rcl} \delta ^{(2)}(\mathbf{r}_{ 1} -\mathbf{r}_{2})\mathbf{I}& =& \int _{\mathrm{S}}\mathbf{K}^{-1}(\mathbf{r}_{ 1},\mathbf{r}_{3}) \cdot \mathbf{K}(\mathbf{r}_{3},\mathbf{r}_{2})d\mathbf{r}_{3} \\ & =& \int _{\mathrm{S}}\mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{3}) \cdot \mathbf{K}^{-1}(\mathbf{r}_{ 3},\mathbf{r}_{2})d\mathbf{r}_{3}{}\end{array}$$
(6.156)

with δ (2)(r) a two-dimensional Dirac delta function and with I the unit tensor as in Eqs. (6.59), then the formal solution of Eq. (6.154) is obtained as

$$\displaystyle{ \mathbf{f}(\mathbf{r}_{1}) = 8\pi \eta _{0}\int _{\mathrm{S}}\mathbf{K}^{-1}(\mathbf{r}_{ 1},\mathbf{r}_{2}) \cdot {\bigl [\mathbf{U}_{\mathrm{c}} +\boldsymbol{ \Omega } \times \mathbf{r}_{2} -\mathbf{v}^{0}(\mathbf{r}_{ 2})\bigr ]}d\mathbf{r}_{2}\,. }$$
(6.157)

We first put v 0 = 0 to consider the translational and rotatory friction tensors of the body, which we denote by \(\boldsymbol{\Xi }\) and \(\boldsymbol{\Xi }_{\mathrm{c,r}}\), respectively. The total frictional force F and torque T c (about the center of mass) exerted by the body on the solvent are then given by

$$\displaystyle\begin{array}{rcl} \mathbf{F} =\boldsymbol{ \Xi } \cdot \mathbf{U}_{\mathrm{c}} =\int _{\mathrm{S}}\mathbf{f}(\mathbf{r})d\mathbf{r}\,,& &{}\end{array}$$
(6.158)
$$\displaystyle\begin{array}{rcl} \mathbf{T}_{\mathrm{c}} =\boldsymbol{ \Xi }_{\mathrm{c,r}} \cdot \boldsymbol{ \Omega } =\int _{\mathrm{S}}\mathbf{r} \times \mathbf{f}(\mathbf{r})d\mathbf{r} =\int _{\mathrm{S}}\mathbf{B}_{\mathrm{c}}(\mathbf{r})^{T} \cdot \mathbf{f}(\mathbf{r})d\mathbf{r}\,,& &{}\end{array}$$
(6.159)

where the superscript T indicates the transpose and the tensor B c is given by

$$\displaystyle{ \mathbf{B}_{\mathrm{c}}(\mathbf{r}) = \left (\begin{array}{*{10}c} 0 & r_{3} & -r_{2} \\ -r_{3} & 0 & r_{1} \\ r_{2} & -r_{1} & 0\\ \end{array} \right ) }$$
(6.160)

with \(\mathbf{r} = r_{1}\mathbf{e}_{1} + r_{2}\mathbf{e}_{2} + r_{3}\mathbf{e}_{3}\). If we substitute Eq. (6.157) into Eqs. (6.158) and (6.159), we find

$$\displaystyle{ \boldsymbol{\Xi } = 8\pi \eta _{0}\int _{\mathrm{S}}\int _{\mathrm{S}}\mathbf{K}^{-1}(\mathbf{r}_{ 1},\mathbf{r}_{2})d\mathbf{r}_{1}d\mathbf{r}_{2} = 8\pi \eta _{0}\int _{\mathrm{S}}\boldsymbol{\Psi }_{1}(\mathbf{r})d\mathbf{r}\,, }$$
(6.161)
$$\displaystyle\begin{array}{rcl} \boldsymbol{\Xi }_{\mathrm{c,r}}& =& 8\pi \eta _{0}\int _{\mathrm{S}}\int _{\mathrm{S}}\mathbf{B}_{\mathrm{c}}(\mathbf{r}_{1})^{T} \cdot \mathbf{K}^{-1}(\mathbf{r}_{ 1},\mathbf{r}_{2}) \cdot \mathbf{B}_{\mathrm{c}}(\mathbf{r}_{2})d\mathbf{r}_{1}d\mathbf{r}_{2} \\ & =& 8\pi \eta _{0}\int _{\mathrm{S}}\mathbf{B}_{\mathrm{c}}(\mathbf{r})^{T} \cdot \boldsymbol{\Psi }_{ 2}(\mathbf{r})d\mathbf{r}\,, {}\end{array}$$
(6.162)

where the tensors \(\boldsymbol{\Psi }_{1}\) and \(\boldsymbol{\Psi }_{2}\) are the solutions of the integral equations

$$\displaystyle\begin{array}{rcl} \int _{\mathrm{S}}\mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{2}) \cdot \boldsymbol{\Psi }_{1}(\mathbf{r}_{2})d\mathbf{r}_{2} = \mathbf{I}\,,& &{}\end{array}$$
(6.163)
$$\displaystyle\begin{array}{rcl} \int _{\mathrm{S}}\mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{2}) \cdot \boldsymbol{\Psi }_{2}(\mathbf{r}_{2})d\mathbf{r}_{2} = \mathbf{B}_{\mathrm{c}}(\mathbf{r}_{1})\,.& &{}\end{array}$$
(6.164)

Now, if we take e 3 along the axis of revolution of the body, then \(\boldsymbol{\Xi }\) and \(\boldsymbol{\Xi }_{\mathrm{c,r}}\) and hence the translational diffusion tensor \(\mathbf{D}_{\mathrm{c}} = k_{\mathrm{B}}T\boldsymbol{\Xi }^{-1}\) (of the center of mass) and rotatory diffusion tensor \(\mathbf{D}_{\mathrm{r}} = k_{\mathrm{B}}T\boldsymbol{\Xi }_{\mathrm{c,r}}^{\ -1}\) are diagonalized. We denote their principal values by \(\Xi _{j}\), \(\Xi _{\mathrm{r},j}\), D j , and D r, j , respectively, so that

$$\displaystyle\begin{array}{rcl} D_{j} = \frac{k_{\mathrm{B}}T} {\Xi _{j}} \,,& &{}\end{array}$$
(6.165)
$$\displaystyle\begin{array}{rcl} D_{\mathrm{r},j} = \frac{k_{\mathrm{B}}T} {\Xi _{\mathrm{r},j}} & &{}\end{array}$$
(6.166)

with D 1 = D 2 and \(D_{\mathrm{r},1} = D_{\mathrm{r},2}\) (and \(\Xi _{1} = \Xi _{2}\) and \(\Xi _{\mathrm{r},1} = \Xi _{\mathrm{r},2}\)). The mean translational diffusion coefficient D averaged over the orientation of the body is given by

$$\displaystyle{ D = \frac{1} {3}\mathrm{Tr}\,\mathbf{D}_{\mathrm{c}}^{(\mathrm{e})} = \frac{1} {3}\mathrm{Tr}\,\mathbf{D}_{\mathrm{c}} = \frac{1} {3}k_{\mathrm{B}}T(2\Xi _{1}^{\ -1} + \Xi _{ 3}^{\ -1})\,. }$$
(6.167)

Next we consider the intrinsic viscosity [η]. The unperturbed flow field v 0 given by Eq. (6.65) may be expressed in the molecular coordinate system as follows,

$$\displaystyle{ \mathbf{v}^{0}(\mathbf{r}) =\epsilon _{ 0}\mathbf{A} \cdot \mathbf{e}_{x}\mathbf{e}_{y} \cdot \mathbf{A}^{T} \cdot \mathbf{r}\,. }$$
(6.168)

We may then put \(\mathbf{U}_{\mathrm{c}} = \mathbf{0}\), so that

$$\displaystyle{ \boldsymbol{\Omega } \times \mathbf{r} -\mathbf{v}^{0}(\mathbf{r}) = -\epsilon _{ 0}\mathbf{m} \cdot \mathbf{r} }$$
(6.169)

with

$$\displaystyle{ \mathbf{m} = \frac{1} {2}\mathbf{A} \cdot (\mathbf{e}_{x}\mathbf{e}_{y} + \mathbf{e}_{y}\mathbf{e}_{x}) \cdot \mathbf{A}^{T}\,, }$$
(6.170)

since the body rotates about the z axis with the angular velocity \(-\epsilon _{0}/2\) in the limit of ε 0 = 0 [27]. Thus Eq. (6.157) becomes

$$\displaystyle{ \mathbf{f}(\mathbf{r}_{1}) = -8\pi \eta _{0}\epsilon _{0}\int _{\mathrm{S}}\mathbf{K}^{-1}(\mathbf{r}_{ 1},\mathbf{r}_{2}) \cdot \mathbf{m} \cdot \mathbf{r}_{2}d\mathbf{r}_{2}\,. }$$
(6.171)

As shown in Appendix 2, [η] of the body may be expressed in terms of the surface integral as

$$\displaystyle{ [\eta ] = -\frac{N_{\mathrm{A}}} {M\eta _{0}\epsilon _{0}}\int _{\mathrm{S}}\mathbf{e}_{x} \cdot {\bigl \langle\mathbf{A}^{T} \cdot \mathbf{f}(\mathbf{r})\mathbf{r} \cdot \mathbf{A}\bigr \rangle} \cdot \mathbf{e}_{ y}d\mathbf{r} = \frac{8\pi N_{\mathrm{A}}} {M} \int _{\mathrm{S}}\Psi (\mathbf{r})d\mathbf{r} }$$
(6.172)

with

$$\displaystyle{ \Psi (\mathbf{r}_{1}) =\int _{\mathrm{S}}\mathbf{e}_{x} \cdot {\bigl \langle\mathbf{A}^{T} \cdot \mathbf{K}^{-1}(\mathbf{r}_{ 1},\mathbf{r}_{2}) \cdot \mathbf{m} \cdot \mathbf{r}_{2}\mathbf{r}_{1} \cdot \mathbf{A}\bigr \rangle} \cdot \mathbf{e}_{y}d\mathbf{r}_{2}\,. }$$
(6.173)

The orientational average in Eq. (6.173) may be evaluated by expanding the matrices A and m in terms of the Wigner functions \(\mathcal{D}_{l}^{mj}(\alpha,\beta,\gamma )\) [89]. The function \(\Psi\) may then be expressed as

$$\displaystyle\begin{array}{rcl} \Psi (\mathbf{r})& =& \frac{1} {3}(-\psi _{11}r_{1} -\psi _{12}r_{2} + 2\psi _{13}r_{3}) + (\psi _{22}r_{3} +\psi _{23}r_{2}) + (\psi _{31}r_{3} +\psi _{33}r_{1}) \\ & & +(\psi _{41}r_{2} +\psi _{42}r_{1}) + (\psi _{51}r_{1} -\psi _{52}r_{2})\,, {}\end{array}$$
(6.174)

where r = (r 1, r 2, \(r_{3}\)), and the vectors \(\boldsymbol{\psi }_{j} = (\psi _{j1}\), ψ j2, ψ j3) (j = 1–5) are the solutions of the integral equations

$$\displaystyle{ \int _{\mathrm{S}}\mathbf{K}(\mathbf{r}_{1},\mathbf{r}_{2}) \cdot \boldsymbol{\psi }_{j}(\mathbf{r}_{2})d\mathbf{r}_{2} = \frac{1} {5}\mathbf{m}_{j} \cdot \mathbf{r}_{1} }$$
(6.175)

with

$$\displaystyle\begin{array}{rcl} & & \qquad \mathbf{m}_{1} = -\mathbf{e}_{1}\mathbf{e}_{1} -\mathbf{e}_{2}\mathbf{e}_{2} + 2\mathbf{e}_{3}\mathbf{e}_{3}\,, \\ & & \mathbf{m}_{2} = \mathbf{e}_{2}\mathbf{e}_{3} + \mathbf{e}_{3}\mathbf{e}_{2}\,,\ \ \ \mathbf{m}_{3} = \mathbf{e}_{3}\mathbf{e}_{1} + \mathbf{e}_{1}\mathbf{e}_{3}\,, \\ & & \mathbf{m}_{4} = \mathbf{e}_{1}\mathbf{e}_{2} + \mathbf{e}_{2}\mathbf{e}_{1}\,,\ \ \ \mathbf{m}_{5} = \mathbf{e}_{1}\mathbf{e}_{1} -\mathbf{e}_{2}\mathbf{e}_{2}\,,{}\end{array}$$
(6.176)

We find exact numerical solutions of all the integral equations above for small \(p = L/d\) and also solutions in the OB approximation (with the non-preaveraged Oseen tensor) for large p. We first consider the latter. General expressions for the transport coefficients of the spheroid-cylinder in the OB approximation may be derived by replacing the integrals over the surface by those over the contour distance s (\(-L/2 \leq s \leq L/2\)), where \(\boldsymbol{\Psi }_{1}\), \(\boldsymbol{\Psi }_{2}\), and \(\boldsymbol{\psi }_{j}\) (j = 1–5) are then functions of s with \(r_{1} = r_{2} = 0\) and r 3 = s.

Then the function f D defined by Eq. (6.37) may be expressed as [27]

$$\displaystyle{ f_{D} = \frac{1} {4}{\bigl (2F_{1}^{\ -1} + F_{ 2}^{\ -1}\bigr )}\,, }$$
(6.177)

where

$$\displaystyle{ F_{j} =\int _{ -1}^{1}\Psi _{ 1j}(x)dx }$$
(6.178)

with \(x = 2s/L\). In Eq. (6.178), \(\Psi _{1j}\) (j = 1, 2) are the solutions of the integral equations

$$\displaystyle{ \int _{-1}^{1}K_{ j}(x_{1},x_{2})\Psi _{1j}(x_{2})dx_{2} = 1\,, }$$
(6.179)

where \(K_{j}(x_{1},x_{2})\) (j = 1, 2) are given by

$$\displaystyle\begin{array}{rcl} K_{1}& =& \frac{2p^{2}(x_{1} - x_{2})^{2} + 3[1 - h(x_{1})]} {d[p^{2}(x_{1} - x_{2})^{2} + 1 - h(x_{1})]^{3/2}}\,,{}\end{array}$$
(6.180)
$$\displaystyle\begin{array}{rcl} K_{2}& =& \frac{4p^{2}(x_{1} - x_{2})^{2} + 2[1 - h(x_{1})]} {d[p^{2}(x_{1} - x_{2})^{2} + 1 - h(x_{1})]^{3/2}}\,,{}\end{array}$$
(6.181)

with \(p = L/d\) and with

$$\displaystyle\begin{array}{rcl} h(x)& =& 0\ \ \ \mathrm{for}\ 0 \leq \vert x\vert <1 - \frac{\epsilon } {p} \\ & =& {\biggl [1 -\frac{p(1 -\vert x\vert )} {\epsilon } \biggr ]}^{2}\ \ \ \mathrm{for}\ 1 - \frac{\epsilon } {p} <\vert x\vert \leq 1\,.{}\end{array}$$
(6.182)

In the limit of \(p \rightarrow \infty\) (with \(h =\epsilon = 0\)), Eq. (6.177) becomes the second line of Eqs. (6.49) [24] if we find the asymptotic solution of Eq. (6.179) by a Legendre polynomial expansion method [20].

The rotatory diffusion coefficient \(D_{\mathrm{r},1}\) and intrinsic viscosity [η] may be expressed as [33]

$$\displaystyle\begin{array}{rcl} D_{\mathrm{r},1} = \frac{3k_{\mathrm{B}}T} {\pi \eta _{0}L^{3}F_{\mathrm{r}}}\,,& &{}\end{array}$$
(6.183)
$$\displaystyle\begin{array}{rcl} [\eta ] = \frac{\pi N_{\mathrm{A}}L^{3}} {90M} (3F_{\mathrm{r}} + F_{\eta _{\infty }})\,,& &{}\end{array}$$
(6.184)

where

$$\displaystyle\begin{array}{rcl} F_{\mathrm{r}} = 3\int _{-1}^{1}x\Psi _{ 21}(x)dx\,,& &{}\end{array}$$
(6.185)
$$\displaystyle\begin{array}{rcl} F_{\eta _{\infty }} = 6\int _{-1}^{1}x\Psi _{ 22}(x)dx& &{}\end{array}$$
(6.186)

with \(\Psi _{2j}\) (j = 1, 2) the solutions of the integral equations

$$\displaystyle{ \int _{-1}^{1}K_{ j}(x_{1},x_{2})\Psi _{2j}(x_{2})dx_{2} = x_{1}\,. }$$
(6.187)

In the OB approximation we have \(\Xi _{\mathrm{r},3} = 0\) from Eqs. (6.162) since \(r_{1} = r_{2} = 0\), so that the rotation of the cylinder about its axis cannot be considered. In the limit of \(p \rightarrow \infty\), we find [33] by the Legendre polynomial expansion method

$$\displaystyle\begin{array}{rcl} F_{\mathrm{r}}^{-1} =\ln p + 2\ln 2 -\frac{11} {6} + \mathcal{O}(p^{-1})\,,& &{}\end{array}$$
(6.188)
$$\displaystyle\begin{array}{rcl} F_{\eta _{\infty }}^{\ -1} =\ln p + 2\ln 2 -\frac{17} {6} + \mathcal{O}(p^{-1})\,,& &{}\end{array}$$
(6.189)

so that Eq. (6.184) becomes Eq. (6.86).

In the following, we complete expressions for f D , D r, 1, and [η].

  1. (a)

    Translational Diffusion Coefficient

As mentioned above and in Sect. 6.3.1, the OB approximation along with the preaveraging of the Oseen tensor and with the KR approximation [3] as in Eq. (6.35) can give the correct asymptotic result for f D as given by Eq. (6.49). We therefore evaluate f D in this way from

$$\displaystyle{ f_{D} = \frac{1} {2}L\int _{-1}^{1}\int _{ -1}^{1}K(x_{ 1},x_{2})dx_{1}dx_{2} }$$
(6.190)

with

$$\displaystyle{ K(x_{1},x_{2}) ={\bigl \{ (x_{1} - x_{2})^{2} +{\bigl [ 1 - h(x_{ 1})\bigr ]}p^{-2}\bigr \}}^{-1/2}\,. }$$
(6.191)

The result reads

$$\displaystyle\begin{array}{rcl} f_{D}& =& \sinh ^{-1}(2p-\epsilon ) - \frac{\epsilon } {p}\sinh ^{-1}\epsilon - \frac{1} {2p}{\bigl \{{\bigl [(2p-\epsilon )^{2} + 1\bigr ]}^{1/2} - (\epsilon ^{2} + 1)^{1/2}\bigr \}} \\ & & - \frac{\epsilon } {2p}\ln \left (2(p-\epsilon ){\bigl \{2p -\epsilon +{\bigl [(2p-\epsilon )^{2} + 1\bigr ]}^{1/2}\bigr \}} + 1\right ) \\ & & + \frac{\epsilon } {2p}\ln {\bigl [4p(p-\epsilon ) + 1\bigr ]} + f'_{D} {}\end{array}$$
(6.192)

with

$$\displaystyle\begin{array}{rcl} f'_{D}& =& \frac{\epsilon } {2p(\epsilon ^{2} - 1)^{1/2}}\biggl \{\epsilon \ln \bigl [\epsilon ^{2} + (\epsilon ^{4} - 1)^{1/2}\bigr ] + (2p-\epsilon ) \\ & & \times \,\ln \biggl |\frac{\epsilon (2p-\epsilon ) - (\epsilon ^{2} - 1)^{1/2}[(2p-\epsilon )^{2} + 1]^{1/2}} {2p[(\epsilon ^{2} - 1)^{1/2}-\epsilon ] + 1} \biggr |\biggr \}\ \ \ \mathrm{for}\ \epsilon> 1 \\ & =& \frac{1} {2p}\bigl \{2p -\bigl [ (2p - 1)^{2} + 1\bigr ]^{1/2} + \sqrt{2}\bigr \}\ \ \ \mathrm{for}\ \epsilon = 1 \\ & =& \frac{\epsilon } {2p(1 -\epsilon ^{2})^{1/2}}\biggl [2\epsilon \tan ^{-1}\biggl (\frac{1 -\epsilon ^{2}} {1 +\epsilon ^{2}}\biggr )^{1/2} + (2p-\epsilon ) \\ & & \times \biggl \{\sin ^{-1} \frac{\epsilon (2p-\epsilon )} {[4p(p-\epsilon ) + 1]^{1/2}} +\sin ^{-1} \frac{1 - 2p\epsilon } {[4p(p-\epsilon ) + 1]^{1/2}}\biggr \}\biggr]\ \ \ \mathrm{for}\ \epsilon <1\,.{}\end{array}$$
(6.193)

Equation (6.192) with Eqs. (6.193) is the desired result and reduces to the result of Norisuye et al. [90] when ε = 1 (prolate spherocylinder), and to the well-known result [91, 92] for the spheroid when ε = p. The latter, which we denote by \(f_{D,(\mathrm{SD})}\), is given by

$$\displaystyle{ f_{D,(\mathrm{SD})} =\epsilon F(\epsilon ) }$$
(6.194)

with

$$\displaystyle\begin{array}{rcl} F(\epsilon )& =& \frac{1} {(\epsilon ^{2} - 1)^{1/2}}\cosh ^{-1}\epsilon \ \ \ \mathrm{for}\ \epsilon> 1 \\ & =& 1\ \ \ \mathrm{for}\ \epsilon = 1 \\ & =& \frac{1} {(1 -\epsilon ^{2})^{1/2}}\cos ^{-1}\epsilon \ \ \ \mathrm{for}\ \epsilon <1\,.{}\end{array}$$
(6.195)

Figure 6.17 shows double-logarithmic plots of f D against p. The solid and dot-dashed curves represent the values calculated from Eq. (6.192) for the spheroid-cylinders with ε = 0. 5 and 1.0 and from Eq. (6.194) for the spheroid (ε = p), respectively, and the unfilled (εp) and filled (ε = p) circles represent the exact numerical solutions. The former values are seen to agree well with the latter. Thus we may adopt Eq. (6.192) with Eqs. (6.193) as a useful interpolation formula for f D for the spheroid-cylinder. When ε = 0, Eq. (6.192) gives the f D, (R) from Eq. (6.39), and the values for this limiting case are represented by the dotted curve. It is seen that the end effects may be ignored for \(p \gtrsim 5\). For comparison, the values calculated from the Broersma equation [18], which are not very different from those from his new version [93], are represented by the dashed curve B. Indeed, his solution of the integral equation is not exact, although asymptotically correct in the limit of \(p \rightarrow \infty\).

Fig. 6.17
figure 17

Double-logarithmic plots of f D against p. The solid, dot-dashed, and dotted curves represent the values calculated from Eq. (6.192) for the spheroid-cylinders, spheroid (ε = p), and cylinder (ε = 0), respectively, and the unfilled and filled circles represent the exact numerical solutions. The dashed curve B represents the values by Broersma [18] for cylinders

  1. (b)

    Rotatory Diffusion Coefficient

We construct an interpolation formula for D r, 1 on the basis of the exact numerical solutions and the OB asymptotic solution above. The result for \(F_{\mathrm{r}} = F_{\mathrm{r}}(p,\epsilon )\) in Eq. (6.183) reads

$$\displaystyle{ F_{\mathrm{r}}(p,\epsilon )^{-1} =\ln p + 2\ln 2 -\frac{11} {6} + \frac{a_{\mathrm{r}0}(\epsilon )} {\ln (1 + p)} +\sum _{ i=1}^{6}a_{\mathrm{ r}i}(\epsilon )\,p^{-i/4} }$$
(6.196)

with

$$\displaystyle\begin{array}{rcl} & & a_{\mathrm{r}0}(\epsilon ) ={\bigl [\ln (1+\epsilon )\bigr ]}{\biggl [f_{\mathrm{r}}(\epsilon )^{-1} -\ln \epsilon -2\ln 2 + \frac{11} {6} -\sum _{i=1}^{6}a_{\mathrm{ r}i}(\epsilon )\,\epsilon ^{-i/4}\biggr ]}\,,{}\end{array}$$
(6.197)
$$\displaystyle\begin{array}{rcl} a_{\mathrm{r}i}(\epsilon ) =\sum _{ j=0}^{2}a_{\mathrm{ r}ij}\,\epsilon ^{j}\,,& &{}\end{array}$$
(6.198)
$$\displaystyle\begin{array}{rcl} f_{\mathrm{r}}(\epsilon )^{-1} = F_{\mathrm{ r}}(\epsilon,\epsilon )^{-1} = \frac{\pi \eta _{0}d^{3}\epsilon ^{3}} {3k_{\mathrm{B}}T}D_{\mathrm{r},1,\mathrm{(SD)}}\,,& &{}\end{array}$$
(6.199)

where a rij are numerical constants and their values are given in Table 6.7; and \(D_{\mathrm{r},1,(\mathrm{SD})}\) is the rotatory diffusion coefficient D r, 1 of the spheroid and is given by [91, 92]

$$\displaystyle\begin{array}{rcl} \frac{\pi \eta _{0}d^{3}D_{\mathrm{r},1,(\mathrm{S}D)}} {k_{\mathrm{B}}T} & =& \frac{3} {2(\epsilon ^{4} - 1)}{\bigl [(2\epsilon ^{2} - 1)F -\epsilon \bigr ]}\ \ \ \mathrm{ for}\ \epsilon \neq 1 \\ & =& 1\ \ \ \mathrm{for}\ \epsilon = 1 {}\end{array}$$
(6.200)

with F being given by Eqs. (6.195). Note that at ε = p, Eq. (6.183) with Eq. (6.196) gives the exact solution given by Eqs. (6.200) for the spheroid. The range of application of Eq. (6.196) is limited to \(0.6 \lesssim \epsilon \lesssim 1.3\). We note that \(D_{\mathrm{r},3,(\mathrm{SD})}\) is given by [91, 92]

$$\displaystyle\begin{array}{rcl} \frac{\pi \eta _{0}d^{3}D_{\mathrm{r},3,(\mathrm{SD})}} {k_{\mathrm{B}}T} & =& \frac{3} {2(\epsilon ^{2} - 1)}(\epsilon -F)\ \ \ \ \mathrm{for}\ \epsilon \neq 1 \\ & =& 1\ \ \ \mathrm{for}\ \epsilon = 1\,, {}\end{array}$$
(6.201)
Table 6.7 Values of \(a_{\mathrm{r}ij}\) in Eq. (6.198)

Figure 6.18 shows double-logarithmic plots of \(D_{\mathrm{r},(\mathrm{S})}/D_{\mathrm{r},1}\) against p, where D r, (S) is the rotatory diffusion coefficient of the Stokes sphere and is equal to \(D_{\mathrm{r},1,(\mathrm{SD})}\) with ε = 1. The solid and dot-dashed curves represent the values calculated from Eq. (6.183) with Eq. (6.196) for the spheroid-cylinders with ε = 0. 63 and 1.0 and from Eqs. (6.200) for the spheroid (ε = p), respectively, and the unfilled (εp) and filled (ε = p) circles represent the exact numerical solutions. It is seen that the former values agree well with the latter, and that the end effects are rather small even for small p. In the figure are also shown the values calculated from the Broersma equation [18], which is also correct only for large p. (Note that they are not very different from the values from his new version [93].)

  1. (c)

    Intrinsic Viscosity

Fig. 6.18
figure 18

Double-logarithmic plots of \(D_{\mathrm{r},(\mathrm{S})}/D_{\mathrm{r},1}\) against p (see the text). The solid and dot-dashed curves represent the values calculated from Eq. (6.183) with Eq. (6.196) for the spheroid-cylinders and spheroid (ε = p), respectively, and the unfilled and filled circles represent the exact numerical solutions. The dashed curve B represents the values by Broersma [18] for cylinders

As in the case of D r, 1, we construct an interpolation formula for [η] on the basis of the exact numerical solutions and the OB asymptotic solution above (with the nonpreaveraged Oseen tensor). The result reads

$$\displaystyle{ [\eta ] = \frac{2\pi N_{\mathrm{A}}L^{3}} {45M} F_{\eta }(p,\epsilon ) }$$
(6.202)

with

$$\displaystyle\begin{array}{rcl} & & F_{\eta }(p,\epsilon )^{-1} =\ln p + 2\ln 2 -\frac{25} {12} + \frac{a_{\eta 0}(\epsilon )} {\ln (1 + p)} +\sum _{ i=1}^{5}a_{\eta i}(\epsilon )\,p^{-i/4}\,,{}\end{array}$$
(6.203)
$$\displaystyle\begin{array}{rcl} & & a_{\eta 0}(\epsilon ) ={\bigl [\ln (1+\epsilon )\bigr ]}{\biggl [f_{\eta }(\epsilon )^{-1} -\ln \epsilon -2\ln 2 + \frac{25} {12} -\sum _{i=1}^{5}a_{\eta i}(\epsilon )\,\epsilon ^{-i/4}\biggr ]}\,,{}\end{array}$$
(6.204)
$$\displaystyle\begin{array}{rcl} a_{\eta i}(\epsilon ) =\sum _{ j=0}^{2}a_{\eta ij}\,\epsilon ^{j}\,,& &{}\end{array}$$
(6.205)
$$\displaystyle\begin{array}{rcl} f_{\eta }(\epsilon ) = F_{\eta }(\epsilon,\epsilon ) = \frac{45M} {2\pi N_{\mathrm{A}}d^{3}\epsilon ^{3}}[\eta ]_{(\mathrm{SD})}\,,& &{}\end{array}$$
(6.206)

where a η i j are numerical constants and their values are given in Table 6.8; and [η](SD) is the intrinsic viscosity of the spheroid and is given by [94, 95]

$$\displaystyle\begin{array}{rcl} [\eta ]_{\mathrm{(SD)}}& =& \frac{\pi N_{\mathrm{A}}d^{3}} {30M} \epsilon (\epsilon ^{2} - 1)^{2}\biggl \{ \frac{2[-(4\epsilon ^{2} - 1)F + 2\epsilon ^{3}+\epsilon ]} {3\epsilon (3F + 2\epsilon ^{3} - 5\epsilon )[(2\epsilon ^{2} + 1)F - 3\epsilon ]} \\ & & + \frac{28} {3\epsilon (3F + 2\epsilon ^{3} - 5\epsilon )} + \frac{4} {(\epsilon ^{2} + 1)(-3\epsilon F +\epsilon ^{2} + 2)} \\ & & + \frac{2(\epsilon ^{2} - 1)} {\epsilon (\epsilon ^{2} + 1)[(2\epsilon ^{2} - 1)F-\epsilon ]}\biggr \}\ \ \ \mathrm{for}\ \epsilon \neq 1 \\ & =& \frac{5\pi N_{\mathrm{A}}d^{3}} {12M} \ \ \ \mathrm{for}\ \epsilon = 1 {}\end{array}$$
(6.207)

with F being given by Eqs. (6.195). Note that at ε = p, Eq. (6.202) gives the exact solution given by Eqs. (6.207) for the spheroid. We also note that the derivation of the first of Eqs. (6.207) by Simha [94] is not correct, although his result happens to be correct, as shown by Saito [95], the second being originally due to Einstein [1, 96]. The range of application of Eq. (6.202) is limited to \(0.6 \lesssim \epsilon \lesssim 1.3\).

Table 6.8 Values of a η i j in Eq. (6.A.52)

Figure 6.19 shows double-logarithmic plots of M[η]∕M 0[η]E p against p, where [η]E is given by Eq. (6.99). The solid and dot-dashed curves represent the values calculated from Eq. (6.202) for the spheroid-cylinders with the indicated values of ε and from Eqs. (6.207) for the spheroid (ε = p), respectively, and the unfilled (εp) and filled (ε = p) circles represent the exact numerical solutions. It is seen that the end effects on [η] are more remarkable than those on D and D r, 1.

Fig. 6.19
figure 19

Double-logarithmic plots of \(M[\eta ]/M_{0}[\eta ]_{\mathrm{E}}p\) against p (see the text). The solid and dot-dashed curves represent the values calculated from Eq. (6.202) for the spheroid-cylinders and spheroid (ε = p), respectively, and the unfilled and filled circles represent the exact numerical solutions

Finally, we note that in the case of cylinders we have found that the values of [η] with the nonpreaveraged and preaveraged Oseen tensors agree with each other to within 1 % for \(60 \lesssim p \lesssim 150\) [27]. This fact has been used in the construction of the interpolation formula for \([\eta ]_{\mathrm{R}}\) given by Eq. (6.92).

Appendix 2: Excess Stress Tensor for the Touched-Bead Model

In this appendix we derive an expression for the excess stress tensor due to an addition of a single touched-bead (or generally subbody) model chain to an incompressible fluid with viscosity coefficient η 0 [38]. In the unperturbed flow field v 0 given by Eq. (6.65) with ε 0 the velocity gradient, the intrinsic viscosity [η] may be written in the form

$$\displaystyle{ [\eta ] = \frac{N_{\mathrm{A}}} {M\eta _{0}\epsilon _{0}}\langle \sigma '_{xy}\rangle = \frac{N_{\mathrm{A}}} {M\eta _{0}\epsilon _{0}}\mathbf{m}:\langle \boldsymbol{\sigma } '\rangle \,, }$$
(6.208)

where m is given by Eq. (6.67), and \(\sigma '_{xy}\) is the xy component of the excess stress tensor \(\boldsymbol{\sigma }'\) for the single chain.

Now the equation of motion for the (incompressible) fluid in steady flow may be written in the form [1]

$$\displaystyle{ \nabla \cdot \boldsymbol{\sigma } (\mathbf{r}) + \mathbf{f}(\mathbf{r}) = \mathbf{0}\,, }$$
(6.209)

where \(\mathbf{f}(\mathbf{r})\) is the force density due to the external (frictional) force exerted on the fluid (per unit volume) at a point r, and \(\boldsymbol{\sigma }\) is the stress tensor given by

$$\displaystyle{ \boldsymbol{\sigma }(\mathbf{r}) = -p(\mathbf{r})\mathbf{I} +\eta _{0}{\bigl \{\nabla \mathbf{v}(\mathbf{r}) +{\bigl [ \nabla \mathbf{v}(\mathbf{r})\bigr ]}^{T}\bigr \}} }$$
(6.210)

with \(\mathbf{v}(\mathbf{r})\) the fluid velocity, p the pressure, I the unit tensor, and the superscript T indicating the transpose. Note that substitution of Eq. (6.210) with Eq. (6.2) into Eq. (6.209) leads to Eq. (6.1). In the present case of the single chain composed of N beads (subbodies), f(r) is given by

$$\displaystyle{ \mathbf{f}(\mathbf{r}) =\sum _{ j=1}^{N}\int _{ \mathrm{S}_{j}}\delta (\mathbf{r} -\mathbf{r}_{j} -\hat{\mathbf{r}}_{j})\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})d\hat{\mathbf{r}}_{j}\,, }$$
(6.211)

where r j is the vector position of the center of the jth bead, \(\hat{\mathbf{r}}_{j}\) is the radius vector from its center to an arbitrary point on its surface, \(\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})\) is the frictional force exerted by the unit area at \(\hat{\mathbf{r}}_{j}\) on the fluid, and the integration is carried out over its surface (S j ) (see Fig. 6.3).

The stress tensor \(\boldsymbol{\sigma }\) may be written as a sum of the stress tensor \(\boldsymbol{\sigma }_{0}\) of the pure fluid and the excess stress tensor \(\boldsymbol{\sigma }'\) due to the force density f; that is, \(\boldsymbol{\sigma }=\boldsymbol{\sigma } _{0} +\boldsymbol{\sigma } '\). Equation (6.209) may therefore be rewritten as

$$\displaystyle\begin{array}{rcl} \nabla \cdot \boldsymbol{\sigma }_{0} = \mathbf{0}\,,& &{}\end{array}$$
(6.212)
$$\displaystyle\begin{array}{rcl} \nabla \cdot \boldsymbol{\sigma }' + \mathbf{f} = \mathbf{0}\,.& &{}\end{array}$$
(6.213)

We take the Fourier transform of both sides of Eq. (6.213),

$$\displaystyle{ i\mathbf{k} \cdot \tilde{\boldsymbol{\sigma }}'(\mathbf{k}) +\tilde{ \mathbf{f}}(\mathbf{k}) = \mathbf{0}\,, }$$
(6.214)

where

$$\displaystyle{ \tilde{\boldsymbol{\sigma }}'(\mathbf{k}) =\int \boldsymbol{\sigma } '(\mathbf{r})\exp (i\mathbf{k} \cdot \mathbf{r})d\mathbf{r}\,, }$$
(6.215)
$$\displaystyle\begin{array}{rcl} \tilde{\mathbf{f}}(\mathbf{k})& =& \int \mathbf{f}(\mathbf{r})\exp (i\mathbf{k} \cdot \mathbf{r})d\mathbf{r} \\ & =& \sum _{j=1}^{N}\int _{ \mathrm{S}_{j}}\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})\exp {\bigl [i\mathbf{k} \cdot (\mathbf{r}_{j} +\hat{ \mathbf{r}}_{j})\bigr ]}d\hat{\mathbf{r}}_{j}\,.{}\end{array}$$
(6.216)

Let \(\mathbf{R}_{\mathrm{c}}\) be the vector position of the center of mass of the chain, and let S j be the vector distance from it to the center of the jth bead. We have \(\mathbf{r}_{j} = \mathbf{R}_{\mathrm{c}} + \mathbf{S}_{j}\), and the second line of Eqs. (6.216) may be rewritten as

$$\displaystyle\begin{array}{rcl} & & \tilde{\mathbf{f}}(\mathbf{k}) =\exp (i\mathbf{k} \cdot \mathbf{R}_{\mathrm{c}})\sum _{j=1}^{N}\mathbf{F}_{ j} + i\mathbf{k} \cdot \exp (i\mathbf{k} \cdot \mathbf{R}_{\mathrm{(C)}}) \\ & & \qquad \times \sum _{j=1}^{N}\int _{ \mathrm{S}_{j}}{\biggl \{\int _{0}^{1}\exp {\bigl [i\xi \mathbf{k} \cdot (\mathbf{S}_{ j} +\hat{ \mathbf{r}}_{j})\bigr ]}d\xi \biggr \}}(\mathbf{S}_{j} +\hat{ \mathbf{r}}_{j})\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})d\hat{\mathbf{r}}_{j}\,,{}\end{array}$$
(6.217)

where F j is the total frictional force exerted by the jth bead and is given by Eq. (6.27). Under the condition of ordinary viscosity measurements, there is not any external force other than shear flow field, so that the total frictional force (sum of F j ) must vanish. We then obtain from Eqs. (6.214) and (6.217)

$$\displaystyle\begin{array}{rcl} \tilde{\boldsymbol{\sigma }}'(\mathbf{k})& =& -\exp (i\mathbf{k} \cdot \mathbf{R}_{\mathrm{c}})\sum _{j=1}^{N}\int _{ \mathrm{S}_{j}}{\biggl \{\int _{0}^{1}\exp {\bigl [i\xi \mathbf{k} \cdot (\mathbf{S}_{ j} +\hat{ \mathbf{r}}_{j})\bigr ]}d\xi \biggr \}} \\ & & \times (\mathbf{S}_{j} +\hat{ \mathbf{r}}_{j})\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})d\hat{\mathbf{r}}_{j}\,. {}\end{array}$$
(6.218)

Finally, we take the configurational average of both sides of Eq. (6.218), noting that R c is distributed uniformly in the fluid and that the average over R c may be taken independently of the other variables. We then obtain

$$\displaystyle{ \langle \tilde{\boldsymbol{\sigma }}'(\mathbf{k})\rangle = -(2\pi )^{3}\delta (\mathbf{k})\sum _{ j=1}^{N}{\biggl [\langle \mathbf{S}_{ j}\mathbf{F}_{j}\rangle +{\biggl \langle\int _{\mathrm{S}_{j}}\hat{\mathbf{r}}_{j}\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})d\hat{\mathbf{r}}_{j}\biggr \rangle}\biggr ]}\,. }$$
(6.219)

Thus, by Fourier inversion of Eq. (6.219), we obtain

$$\displaystyle{ \langle \boldsymbol{\sigma }'\rangle = -\sum _{j=1}^{N}{\biggl [\langle \mathbf{F}_{ j}\mathbf{S}_{j}\rangle +{\biggl \langle\int _{\mathrm{S}_{i}}\mathbf{f}_{j}(\hat{\mathbf{r}}_{j})\hat{\mathbf{r}}_{j}d\hat{\mathbf{r}}_{j}\biggr \rangle}\biggr ]}\,. }$$
(6.220)

Substitution of Eq. (6.220) into the second of Eqs. (6.208) leads to Eq. (6.102). In the case of a single rigid body, it also reduces to the first of Eqs. (6.172).

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Yamakawa, H., Yoshizaki, T. (2016). Transport Properties. In: Helical Wormlike Chains in Polymer Solutions. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-662-48716-7_6

Download citation

Publish with us

Policies and ethics