Skip to main content
Log in

Vessel radius mapping in an extended model of transverse relaxation

  • Research Article
  • Published:
Magnetic Resonance Materials in Physics, Biology and Medicine Aims and scope Submit manuscript

Abstract

Objectives

Spin dephasing of the local magnetization in blood vessel networks can be described in the static dephasing regime (where diffusion effects may be ignored) by the established model of Yablonskiy and Haacke. However, for small capillary radii, diffusion phenomena for spin-bearing particles are not negligible.

Material and methods

In this work, we include diffusion effects for a set of randomly distributed capillaries and provide analytical expressions for the transverse relaxation times T2* and T2 in the strong collision approximation and the Gaussian approximation that relate MR signal properties with microstructural parameters such as the mean local capillary radius.

Results

Theoretical results are numerically validated with random walk simulations and are used to calculate capillary radius distribution maps for glioblastoma mouse brains at 9.4 T. For representative tumor regions, the capillary maps reveal a relative increase of mean radius for tumor tissue towards healthy brain tissue of \(128 \pm 23 \%\) (p < 0.001).

Conclusion

The presented method may be used to quantify angiogenesis or the effects of antiangiogenic therapy in tumors whose growth is associated with significant microvascular changes.

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

Similar content being viewed by others

Abbreviations

CBV:

Cerebral blood volume

CV:

Coefficient of variation

DSC:

Dynamic susceptibility contrast

GE:

Gradient echo

Hct:

Hematocrit

OEF:

Oxygen extraction fraction

RD:

Radius distribution

RO:

Random orientation

RVM:

Random vessel distribution model

SCA:

Strong collision approximation

StDe:

Standard deviation

SE:

Spin echo

StEr:

Standard error

SVM:

Single-vessel approximation model

WFA:

Weak field approximation

References

  1. Reichenbach JR, Schweser F, Serres B, Deistung A (2015) Quantitative susceptibility mapping: concepts and applications. Clin Neuroradiol 25:225–230

    Article  PubMed  Google Scholar 

  2. Liu S, Buch S, Chen Y, Choi HS, Dai Y, Habib C, Hu J, Jung JY, Luo Y, Utriainen D, Wang M, Wu D, Xia S, Haacke EM (2017) Susceptibility-weighted imaging: current status and future directions. NMR Biomed 30:e3552

    Article  CAS  Google Scholar 

  3. Ogawa S, Lee TM, Kay AR, Tank DW (1990) Brain magnetic resonance imaging with contrast dependent on blood oxygenation. Proc Natl Acad Sci USA 87:9868–9872

    Article  PubMed  CAS  Google Scholar 

  4. Sehgal V, Delproposto Z, Haacke EM, Tong KA, Wycliffe N, Kido DK, Xu Y, Neelavalli J, Haddar D, Reichenbach JR (2005) Clinical applications of neuroimaging with susceptibility-weighted imaging. J Magn Reson Imaging 22:439–450

    Article  PubMed  Google Scholar 

  5. Deistung A, Schäfer A, Schweser F, Biedermann U, Turner R, Reichenbach JR (2013) Toward in vivo histology: A comparison of quantitative susceptibility mapping (QSM) with magnitude-, phase-, and R2*-imaging at ultra-high magnetic field strength. Neuroimage 65:299–314

    Article  PubMed  Google Scholar 

  6. Liu C (2010) Susceptibility tensor imaging. Magn Reson Med 63:1471–1477

    Article  PubMed  PubMed Central  Google Scholar 

  7. Pauling L, Coryell CD (1936) The magnetic properties and structure of hemoglobin, oxyhemoglobin and carbonmonoxyhemoglobin. Proc Natl Acad Sci USA 22:210–216

    Article  PubMed  CAS  Google Scholar 

  8. Meyer EP, Ulmann-Schuler A, Staufenbiel M, Krucker T (2008) Altered morphology and 3D architecture of brain vasculature in a mouse model for Alzheimer’s disease. Proc Natl Acad Sci USA 105:3587–3592

    Article  PubMed  Google Scholar 

  9. Lauwers F, Cassot F, Lauwers-Cances V, Puwanarajah P, Duvernoy H (2008) Morphometry of the human cerebral cortex microcirculation: general characteristics and space-related profiles. Neuroimage 39:936–948

    Article  PubMed  Google Scholar 

  10. Yablonskiy DA, Haacke EM (1994) Theory of NMR signal behavior in magnetically inhomogeneous tissues: the static dephasing regime. Magn Reson Med 32:749–763

    Article  PubMed  CAS  Google Scholar 

  11. Yablonskiy DA, Sukstanskii AL, He X (2013) Blood oxygenation level-dependent (BOLD)-based techniques for the quantification of brain hemodynamic and metabolic properties-theoretical models and experimental approaches. NMR Biomed 26:963–986

    Article  PubMed  CAS  Google Scholar 

  12. Buxton RB (2013) The physics of functional magnetic resonance imaging (fMRI). Rep Prog Phys 76:096601

    Article  PubMed  PubMed Central  Google Scholar 

  13. Grebenkov DS (2007) NMR survey of the reflected Brownian motion. Rev Mod Phys 79:1077–1137

    Article  CAS  Google Scholar 

  14. Torrey HC (1956) Bloch equations with diffusion terms. Phys Rev 104:563–565

    Article  Google Scholar 

  15. Jensen JH, Chandra R (2000) NMR relaxation in tissues with weak magnetic inhomogeneities. Magn Reson Med 44:144–156

    Article  PubMed  CAS  Google Scholar 

  16. Jensen JH, Chandra R, Ramani A, Lu H, Johnson G, Lee SP, Kaczynski K, Helpern JA (2006) Magnetic field correlation imaging. Magn Reson Med 55:1350–1361

    Article  PubMed  Google Scholar 

  17. Sukstanskii AL, Yablonskiy DA (2004) Gaussian approximation in the theory of MR signal formation in the presence of structure-specific magnetic field inhomogeneities. Effects of impermeable susceptibility inclusions. J Magn Reson 167:56–67

    Article  PubMed  CAS  Google Scholar 

  18. Ziener CH, Kampf T, Melkus G, Herold V, Weber T, Reents G, Jakob PM, Bauer WR (2007) Local frequency density of states around field inhomogeneities in magnetic resonance imaging: effects of diffusion. Phys Rev E 76:031915

    Article  CAS  Google Scholar 

  19. Bauer WR, Nadler W, Bock M, Schad LR, Wacker C, Hartlep A, Ertl G (1999) The relationship between the BOLD-induced T2 and T2*: a theoretical approach for the vasculature of myocardium. Magn Reson Med 42:1004–1010

    Article  PubMed  CAS  Google Scholar 

  20. Dickson JD, Ash TW, Williams GB, Sukstanskii AL, Ansorge RE, Yablonskiy DA (2011) Quantitative phenomenological model of the BOLD contrast mechanism. J Magn Reson 212:17–25

    Article  PubMed  CAS  Google Scholar 

  21. Krogh A (1919) The supply of oxygen to the tissues and the regulation of the capillary circulation. J Physiol 52:457–474

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Ziener CH, Kurz FT, Kampf T (2015) Free induction decay caused by a dipole field. Phys Rev E 91:032707

    Article  CAS  Google Scholar 

  23. Kurz FT, Kampf T, Buschle LR, Heiland S, Schlemmer HP, Bendszus M, Ziener CH (2016) CPMG relaxation rate dispersion in dipole fields around capillaries. Magn Reson Imaging 34:875–888

    Article  PubMed  CAS  Google Scholar 

  24. Ziener CH, Kampf T, Reents G, Schlemmer HP, Bauer WR (2012) Spin dephasing in a magnetic dipole field. Phys Rev E 85:051908

    Article  CAS  Google Scholar 

  25. Ziener CH, Kampf T, Melkus G, Jakob PM, Schlemmer HP, Bauer WR (2012) Signal evolution in the local magnetic field of a capillary—analogy to the damped driven harmonic oscillator. Magn Reson Imaging 30:540–553

    Article  PubMed  Google Scholar 

  26. Kurz FT, Buschle LR, Kampf T, Zhang K, Schlemmer HP, Heiland S, Bendszus M, Ziener CH (2016) Spin dephasing in a magnetic dipole field around large capillaries: approximative and exact results. J Magn Reson 273:83–97

    Article  PubMed  CAS  Google Scholar 

  27. Harrison RV, Harel N, Panesar J, Mount RJ (2002) Blood capillary distribution correlates with hemodynamic-based functional imaging in cerebral cortex. Cereb Cortex 12:225–233

    Article  PubMed  Google Scholar 

  28. Bauer WR, Nadler W, Bock M, Schad LR, Wacker C, Hartlep A, Ertl G (1999) Theory of the BOLD effect in the capillary region: an analytical approach for the determination of T2 in the capillary network of myocardium. Magn Reson Med 41:51–62

    Article  PubMed  CAS  Google Scholar 

  29. Breckwoldt MO, Bode J, Kurz FT, Hoffmann A, Ochs K, Ott M, Deumelandt K, Krüwel T, Schwarz D, Fischer M, Helluy X, Milford D, Kirschbaum K, Solecki G, Chiblak S, Abdollahi A, Winkler F, Wick W, Platten M, Heiland S, Bendszus M, Tews B (2016) Correlated magnetic resonance imaging and ultramicroscopy (MR-UM) is a tool kit to assess the dynamics of glioma angiogenesis. eLife 5:e11712

    Article  PubMed  PubMed Central  Google Scholar 

  30. Spees WM, Yablonskiy DA, Oswood MC, Ackerman JJH (2001) Water proton MR properties of human blood at 1.5 Tesla: magnetic susceptibility, T1, T2, T2*, and non-Lorentzian signal behavior. Magn Reson Med 45:533–542

    Article  PubMed  CAS  Google Scholar 

  31. Brown RW, Haacke EM, Cheng YC, Thompson MR, Venkatesan R (2014) Magnetic resonance imaging: physical principles and sequence design. John Wiley & Sons, Hoboken

    Book  Google Scholar 

  32. Boxerman JL, Hamberg LM, Rosen BR, Weisskoff RM (1995) MR contrast due to intravascular magnetic susceptibility perturbations. Magn Reson Med 34:555–566

    Article  PubMed  CAS  Google Scholar 

  33. Purves WK, Orians GH, Sadava D, Heller HC (2003) Life: the science of biology: volume III: plants and animals. Macmillan, Basingstoke

    Google Scholar 

  34. He X, Zhu M, Yablonskiy DA (2008) Validation of oxygen extraction fraction measurement by qBOLD technique. Magn Reson Med 60:882–888

    Article  PubMed  PubMed Central  Google Scholar 

  35. Sukstanskii AL, Yablonskiy DA (2003) Gaussian approximation in the theory of MR signal formation in the presence of structure-specific magnetic field inhomogeneities. J Magn Reson 163:236–247

    Article  PubMed  CAS  Google Scholar 

  36. Ziener CH, Kampf T, Jakob PM, Bauer WR (2010) Diffusion effects on the CPMG relaxation rate in a dipolar field. J Magn Reson 202:38–42

    Article  PubMed  CAS  Google Scholar 

  37. Chugh BP, Lerch JP, Lisa XY, Pienkowski M, Harrison RV, Henkelman RM, Sled JG (2009) Measurement of cerebral blood volume in mouse brain regions using micro-computed tomography. Neuroimage 47:1312–1318

    Article  PubMed  Google Scholar 

  38. Ziener CH, Kampf T, Kurz FT (2015) Diffusion propagators for hindered diffusion in open geometries. Concepts Magn Reson A 44:150–159

    Article  CAS  Google Scholar 

  39. Kurz FT, Kampf T, Buschle LR, Schlemmer HP, Bendszus M, Heiland S, Ziener CH (2016) Generalized moment analysis of magnetic field correlations for accumulations of spherical and cylindrical magnetic pertubers. Front Phys 4:46

    Article  Google Scholar 

  40. Kiselev VG, Posse S (1998) Analytical theory of susceptibility induced NMR signal dephasing in a cerebrovascular network. Phys Rev Lett 81:5696–5699

    Article  CAS  Google Scholar 

  41. Kiselev VG, Posse S (1999) Analytical model of susceptibility-induced MR signal dephasing: effect of diffusion in a microvascular network. Magn Reson Med 41:499–509

    Article  PubMed  CAS  Google Scholar 

  42. Dattagupta S, Blume M (1974) Stochastic theory of line shape. I. Nonsecular effects in the strong-collision model. Phys Rev B 10:4540–4550

    Article  CAS  Google Scholar 

  43. Bauer WR, Nadler W, Bock M, Schad LR, Wacker C, Hartlep A, Ertl G (1999) Theory of coherent and incoherent nuclear spin dephasing in the heart. Phys Rev Lett 83:4215–4218

    Article  CAS  Google Scholar 

  44. Kurz FT, Ziener CH, Rückl M, Hahn A, Sturm VJF, Zhang K, Buschle LR, Bendszus M, Heiland S, Schlemmer HP, Bauer WR, Kampf T (2017) The influence of spatial patterns of capillary networks on transverse relaxation. Magn Reson Imaging 40:31–47

    Article  PubMed  CAS  Google Scholar 

  45. Mooney CZ (1997) Monte Carlo simulation. Sage Publications, Thousand Oaks

    Book  Google Scholar 

  46. Hall MG, Alexander DC (2009) Convergence and parameter choice for Monte-Carlo simulations of diffusion MRI. IEEE Trans Med Imaging 28:1354–1364

    Article  PubMed  Google Scholar 

  47. Knuth DE (2013) Art of computer programming. Addison-Wesley, Boston

    Google Scholar 

  48. von Baumgarten L, Brucker D, Tirniceru A, Kienast Y, Grau S, Burgold S, Herms J, Winkler F (2011) Bevacizumab has differential and dose-dependent effects on glioma blood vessels and tumor cells. Clin Cancer Res 17:6192–6205

    Article  Google Scholar 

  49. Hennig J, Nauerth A, Friedburg H (1986) RARE imaging: a fast imaging method for clinical MR. Magn Reson Med 3:823–833

    Article  PubMed  CAS  Google Scholar 

  50. Stables LA, Kennan RP, Gore JC (1998) Asymmetric spin-echo imaging of magnetically inhomogeneous systems: theory, experiment, and numerical studies. Magn Reson Med 40:432–442

    Article  PubMed  CAS  Google Scholar 

  51. Weisskoff RM, Zuo CS, Boxerman JL, Rosen BR (1994) Microscopic susceptibility variation and transverse relaxation: theory and experiment. Magn Reson Med 31:601–610

    Article  PubMed  CAS  Google Scholar 

  52. Cassot F, Lauwers F, Fouard C, Prohaska S, Lauwers-Cances V (2006) A novel three-dimensional computer-assisted method for a quantitative study of microvascular networks of the human cerebral cortex. Microcirculation 13:1–18

    Article  PubMed  CAS  Google Scholar 

  53. Duvernoy HM, Delon S, Vannson JL (1981) Cortical blood vessels of the human brain. Brain Res Bull 7:519–579

    Article  PubMed  CAS  Google Scholar 

  54. Duvernoy H, Delon S, Vannson JL (1983) The vascularization of the human cerebellar cortex. Brain Res Bull 11:419–480

    Article  PubMed  CAS  Google Scholar 

  55. van Zijl PCM, Eleff SM, Ulatowski JA, Ulug AM, Traystman RJ, Kauppinen RA (1998) Quantitative assessment of blood flow, blood volume and blood oxygenation effects in functional magnetic resonance imaging. Nat Med 4:159–167

    Article  PubMed  Google Scholar 

  56. Martindale J, Kennerley AJ, Johnston D, Zheng Y, Mayhew JE (2008) Theory and generalization of Monte Carlo models of the BOLD signal source. Magn Reson Med 59:607–618

    Article  PubMed  Google Scholar 

  57. Grandin CB, Bol A, Smith AM, Michel C, Cosnard G (2005) Absolute CBF and CBV measurements by MRI bolus tracking before and after acetazolamide challenge: repeatability and comparison with PET in humans. Neuroimage 26:525–535

    PubMed  Google Scholar 

  58. Pathak AP, Kim E, Zhang J, Jones MV (2011) Three-dimensional imaging of the mouse neurovasculature with magnetic resonance microscopy. PLoS ONE 6:e22643

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  59. Guilfoyle DN, Dyakin VV, O’Shea J, Pell GS, Helpern JA (2003) Quantitative measurements of proton spin-lattice (T1) and spin-spin (T2) relaxation times in the mouse brain at 7.0 T. Magn Reson Med 49:576–580

    Article  PubMed  Google Scholar 

  60. Le Bihan D (1995) Molecular diffusion, tissue microdynamics and microstructure. NMR Biomed 8:375–386

    Article  PubMed  Google Scholar 

  61. Troprés I, Lamalle L, Peoc’h M, Farion R, Usson Y, Decorps M, Remy C (2004) In vivo assessment of tumoral angiogenesis. Magn Reson Med 51:533–541

    Article  PubMed  Google Scholar 

  62. Valable S, Lemasson B, Farion R, Beaumont M, Segebarth C, Remy C, Barbier EL (2008) Assessment of blood volume, vessel size, and the expression of angiogenic factors in two rat glioma models: a longitudinal in vivo and ex vivo study. NMR Biomed 21:1043–1056

    Article  PubMed  CAS  Google Scholar 

  63. Ziener CH, Bauer WR, Jakob PM (2005) Transverse relaxation of cells labeled with magnetic nanoparticles. Magn Reson Med 54:702–706

    Article  PubMed  CAS  Google Scholar 

  64. Buschle LR, Kurz FT, Kampf T, Triphan SMF, Schlemmer HP, Ziener CH (2015) Diffusion-mediated dephasing in the dipole field around a single spherical magnetic object. Magn Reson Imaging 33:1126–1145

    Article  PubMed  Google Scholar 

  65. Mueller J, Raisi H, Rausch V, Peccerella T, Simons D, Ziener CH, Schlemmer HP, Seitz HK, Waldburger N, Longerich T, Straub BK, Mueller S (2017) Sensitive and non-invasive assessment of hepatocellular iron using a novel room-temperature susceptometer. J Hepatol 67:535–542

    Article  PubMed  CAS  Google Scholar 

  66. Winkler F, Kienast Y, Fuhrmann M, Von Baumgarten L, Burgold S, Mitteregger G, Kretzschmar H, Herms J (2009) Imaging glioma cell invasion in vivo reveals mechanisms of dissemination and peritumoral angiogenesis. Glia 57:1306–1315

    Article  PubMed  Google Scholar 

  67. Farrar CT, Kamoun WS, Ley CD, Kim YR, Kwon SJ, Dai G, Rosen BR, di Tomaso E, Jain RK, Sorensen AG (2010) In vivo validation of MRI vessel caliber index measurement methods with intravital optical microscopy in a U87 mouse brain tumor model. Neuro Oncol 12(4):341–350

    Article  PubMed  PubMed Central  Google Scholar 

  68. Quail DF, Joyce JA (2013) Microenvironmental regulation of tumor progression and metastasis. Nat Med 19:1432–1437

    Article  CAS  Google Scholar 

  69. Linkous AG, Yazlovitskaya EM (2011) Angiogenesis in glioblastoma multiforme: navigating the maze. Anticancer Agents Med Chem 11:712–718

    Article  PubMed  CAS  Google Scholar 

  70. Cipolla MJ (2009) The cerebral circulation. Morgan & Claypool Life Sciences, San Rafael

    Google Scholar 

  71. Fukunaga M, Li TQ, van Gelderen P, de Zwart JA, Shmueli K, Yao B, Maric Lee J, Aronova MA, Zhang G, Leapman RD, Schenck JF, Merkle H, Duyn JH (2010) Layer-specific variation of iron content in cerebral cortex as a source of MRI contrast. Proc Natl Acad Sci USA 107:3834–3839

    Article  PubMed  Google Scholar 

  72. Hocq A, Brouette N, Saussez S, Luhmer M, Gillis P, Gossuin Y (2009) Variable-field relaxometry of iron-containing human tissues: a preliminary study. Contrast Media Mol Imaging 4:157–164

    Article  PubMed  CAS  Google Scholar 

  73. Yablonskiy DA (1998) Quantitation of intrinsic magnetic susceptibility-related effects in a tissue matrix. Phantom study. Magn Reson Med 39:417–428

    Article  PubMed  CAS  Google Scholar 

  74. Domsch S, Mie MB, Wenz F, Schad LR (2014) Non-invasive multiparametric qBOLD approach for robust mapping of the oxygen extraction fraction. Z Med Phys 24:231–242

    Article  PubMed  Google Scholar 

  75. Christen T, Bolar DS, Zaharchuk G (2013) Imaging brain oxygenation with MRI using blood oxygenation approaches: methods, validation, and clinical applications. AJNR Am J Neuroradiol 34:1113–1123

    Article  PubMed  CAS  Google Scholar 

  76. Troprés I, Grimault S, Vaeth A, Grillon E, Julien C, Payen JF, Lamalle L, Decorps M (2001) Vessel size imaging. Magn Reson Med 45(3):397–408

    Article  PubMed  Google Scholar 

  77. Wen J, Cross AH, Yablonskiy DA (2015) On the role of physiological fluctuations in quantitative gradient echo MRI: implications for GEPCI, QSM, and SWI. Magn Reson Med 73:195–203

    Article  PubMed  Google Scholar 

  78. Ulrich X, Yablonskiy DA (2016) Separation of cellular and BOLD contributions to T2* signal relaxation. Magn Reson Med 75:606–615

    Article  PubMed  Google Scholar 

  79. Kennan RP, Zhong J, Gore JC (1994) Intravascular susceptibility contrast mechanisms in tissues. Magn Reson Med 31:9–21

    Article  PubMed  CAS  Google Scholar 

  80. Buschle LR, Kurz FT, Kampf T, Wagner WL, Duerr J, Stiller W, Konietzke P, Wünnemann F, Mall MA, Wielpütz MO, Schlemmer HP, Ziener CH (2017) Dephasing and diffusion on the alveolar surface. Phys Rev E 95:022415

    Article  PubMed  CAS  Google Scholar 

  81. Emblem KE, Farrar CT, Gerstner ER, Batchelor TT, Borra RJH, Rosen BR, Sorensen AG, Jain RK (2014) Vessel calibre—a potential MRI biomarker of tumour response in clinical trials. Nat Rev Clin Oncol 11:566–584

    Article  PubMed  PubMed Central  Google Scholar 

  82. Emblem KE, Mouridsen K, Bjornerud A, Farrar CT, Jennings D, Borra RJH, Wen PY, Ivy P, Batchelor TT, Rosen BR, Jain RK, Sorensen AG (2013) Vessel architectural imaging identifies cancer patient responders to anti-angiogenic therapy. Nat med 19(9):1178–1183

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  83. Dennie J, Mandeville JB, Boxerman JL, Packard SD, Rosen BR, Weisskoff RM (1998) NMR imaging of changes in vascular morphology due to tumor angiogenesis. Magn Reson Med 40(6):793–799

    Article  PubMed  CAS  Google Scholar 

  84. Troprés I, Pannetier N, Grand S, Lemasson B, Moisan A, Péoc’h M, Rémy C, Barbier EL (2015) Imaging the microvessel caliber and density: principles and applications of microvascular MRI. Magn Reson Med 73:325–341

    Article  PubMed  Google Scholar 

  85. Stadlbauer A, Zimmermann M, Heinz G, Oberndorfer S, Doerfler A, Buchfelder M, Rössler K (2017) Magnetic resonance imaging biomarkers for clinical routine assessment of microvascular architecture in glioma. J Cereb Blood Flow Metab 37:632–643

    Article  PubMed  Google Scholar 

  86. Jochimsen TH, Moller HE (2008) Increasing specificity in functional magnetic resonance imaging by estimation of vessel size based on changes in blood oxygenation. Neuroimage 40:228–236

    Article  PubMed  Google Scholar 

  87. Pannetier NA, Sohlin M, Christen T, Schad L, Schuff N (2014) Numerical modeling of susceptibility-related MR signal dephasing with vessel size measurement: phantom validation at 3T. Magn Reson Med 72:646–658

    Article  PubMed  Google Scholar 

  88. Kording F, Weidensteiner C, Zwick S, Osterberg N, Weyerbock A, Staszewski O, Elverfeldt D, Reichardt W (2014) Simultaneous assessment of vessel size index, relative blood volume, and vessel permeability in a mouse brain tumor model using a combined spin echo gradient echo echo-planar imaging sequence and viable tumor analysis. J Magn Reson Imaging 40:1310–1318

    Article  PubMed  Google Scholar 

  89. Bonekamp D, Mouridsen K, Radbruch A, Kurz FT, Eidel O, Wick A, Schlemmer HP, Wick W, Bendszus M, Østergaard L (2017) Kickingereder P (2017) Assessment of tumor oxygenation and its impact on treatment response in bevacizumab-treated recurrent glioblastoma. J Cereb Blood Flow Metab 37:485–494

    Article  PubMed  Google Scholar 

  90. Ziener CH, Glutsch S, Jakob PM, Bauer WR (2009) Spin dephasing in the dipole field around capillaries and cells: numerical solution. Phys Rev E 80:046701

    Article  CAS  Google Scholar 

  91. Ziener CH, Bauer WR, Jakob PM (2005) Frequency distribution and signal formation around a vessel. Magn Reson Mater Phy 18:225–230

    Article  CAS  Google Scholar 

  92. Ziener CH, Rückl M, Kampf T, Bauer WR, Schlemmer HP (2012) Mathieu functions for purely imaginary parameters. J Comput Appl Math 236:4513–4524

    Article  Google Scholar 

  93. Ziener CH, Kurz FT, Buschle LR, Kampf T (2015) Orthogonality, Lommel integrals and cross product zeros of linear combinations of Bessel functions. SpringerPlus 4:390

    Article  PubMed  PubMed Central  Google Scholar 

  94. Moiseyev N (2011) Non-Hermitian quantum mechanics. Cambridge University Press, Cambridge

    Book  Google Scholar 

  95. Bagarello F, Gazeau JP, Szafraniec FH, Znojil M (2015) Non-selfadjoint operators in quantum physics: mathematical aspects. John Wiley & Sons, Hoboken

    Book  Google Scholar 

  96. Herberthson M, Özarslan E, Knutsson H, Westin CF (2017) Dynamics of local magnetization in the eigenbasis of the Bloch-Torrey operator. J Chem Phys 146:12401

    Article  CAS  Google Scholar 

  97. Seeger A (1979) The determination of the diffusion coefficient of positive muons from the transverse depolarization in type-II superconductors. Phys Lett A 74:259–262

    Article  Google Scholar 

  98. Seeger A (1997) Transverse spin relaxation of spin carriers diffusing in spatially periodic magnetic fields. Hyperfine Interact 105:151–159

    Article  CAS  Google Scholar 

  99. Stoller SD, Happer W, Dyson FJ (1991) Transverse spin relaxation in inhomogeneous magnetic fields. Phys Rev A 44:7459–7477

    Article  PubMed  CAS  Google Scholar 

  100. Bender CM, Brody DC, Jones HF (2003) Must a hamiltonian be hermitian? Am J Phys 71:1095–1102

    Article  CAS  Google Scholar 

  101. Gohberg IC, Krein MG (1969) Introduction to the theory of linear nonselfadjoint operators. American Mathematical Society, Providence

    Google Scholar 

  102. Widder DV (1946) The Laplace transformation. Princeton University Press, Princeton

    Google Scholar 

  103. Kurz FT, Kampf T, Heiland S, Bendszus M, Schlemmer HP, Ziener CH (2014) Theoretical model of the single spin-echo relaxation time for spherical magnetic perturbers. Magn Reson Med 71:1888–1895

    Article  PubMed  Google Scholar 

  104. Kurz FT, Kampf T, Buschle LR, Schlemmer HP, Heiland S, Bendszus M, Ziener CH (2015) Microstructural analysis of peripheral lung tissue through CPMG inter-echo time R2 dispersion. PLoS One 10:e0141894

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Ziener CH, Kampf T, Melkus G, Jakob PM, Bauer WR (2007) Scaling laws for transverse relaxation times. J Magn Reson 184:169–175

    Article  PubMed  CAS  Google Scholar 

  106. Ziener CH, Kampf T, Herold V, Jakob PM, Bauer WR, Nadler W (2008) Frequency autocorrelation function of stochastically fluctuating fields caused by specific magnetic field inhomogeneities. J Chem Phys 129:014507

    Article  PubMed  CAS  Google Scholar 

  107. Bauer WR, Ziener CH, Jakob PM (2005) Non-Gaussian spin dephasing. Phys Rev A 71:053412

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by grants from the Deutsche Forschungsgemeinschaft (contract grant number: DFG ZI 1295/2-1 and DFG KU 3555/1-1). L. R. Buschle was supported by a scholarship of the German Academic Scholarship Foundation (Studienstiftung des deutschen Volkes). F. T. Kurz was supported by a postdoctoral fellowship from the medical faculty of Heidelberg University and the Hoffmann-Klose foundation of Heidelberg University.

Author information

Authors and Affiliations

Authors

Contributions

This work was carried out by the all authors in collaboration. LRB and FTK designed the research; LRB, CHZ, KZ, VJFS, AH, GS and FTK performed the research; LRB contributed numerical tools; and LRB, CHZ, TK, HS, FW, MB, SH, HPS and FTK wrote the paper.

Corresponding author

Correspondence to Felix T. Kurz.

Ethics declarations

Conflict of interest

The authors declare that they have no conflicts of interest.

Ethical approval

All applicable international, national and/or institutional guidelines for the care and use of animals were followed. All procedures performed in studies involving animals were in accordance with the ethical standards of the institution or practice at which the studies were conducted.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 893 KB)

Appendices

Appendix A: Single-vessel approximation model (SVM)

In this appendix, the random vessel distribution model (RVM) is compared to the single-vessel approximation model (SVM). Static dephasing limit, an exact solution of the Bloch-Torrey equation in the SVM and the SCA in the SVM were summarized and compared to the obtained results in the RVM.

Geometrical model

The Bloch-Torrey equation can be solved for uniformly arranged capillaries by applying Krogh’s capillary supply model [21] to reduce the vessel network geometry to one single vessel. In this model, only one single capillary with radius R is considered, which is surrounded by a second concentric cylinder, or Krogh cylinder, that accounts for the surrounding tissue. The outer radius \(R_\text {D}\) is chosen such that the capillary blood volume \(\eta \) corresponds to:

$$\begin{aligned} \frac{N\pi R^2}{V} = \eta = \frac{R^2}{R_\text {D}^2}\,. \end{aligned}$$
(38)

Reflecting boundary conditions are chosen at the surface of the Krogh cylinder (see e.g. Fig. 1 in [90]). A schematic illustration of the SVM is shown in Fig. 1c.

This simplistic geometrical model provides an exact solution of the Bloch-Torrey Eq. (1) for spin dephasing around capillaries, as demonstrated recently (see also below and [22, 24, 26]).

Static dephasing regime

In the static dephasing regime (index 0), where diffusion effects are neglected, the free induction decay in the SVM can be found as the signal around a single blood vessel (see [18] and Eq. (36) in [10]):

$$\begin{aligned} M_{0,\text {SVM}}(\eta ,t)&= \frac{h(\eta \delta \omega t)-\eta h(\delta \omega t)}{1-\eta } = \frac{\eta }{1-\eta } \int \limits _{\eta }^{1} \frac{\mathrm {d}x}{x^2} J_0(x\delta \omega t)\end{aligned}$$
(39)
$$\begin{aligned}&= \frac{2}{\pi } \left[ \int \limits _\eta ^1 \!\!\! \mathrm {d}x \frac{\sin (x\,\delta \omega \,t) \sqrt{x^2 - \eta ^2}}{x^2[1-\eta ]} + \int \limits _1^\infty \!\!\! \mathrm {d}x \frac{\sin (x\,\delta \omega \, t)\left[ \sqrt{x^2-\eta ^2} + \eta \sqrt{x^2-1} \right] [1+\eta ]}{\displaystyle {x^2[1+\eta ^2] + 2\eta \left[ \sqrt{[x^2-\eta ^2][x^2-1]} - \eta \right] }}\right] \,, \end{aligned}$$
(40)

where \(J_0(x\delta \omega t)\) denotes the Bessel function of first kind with an index of zero. The function h(x) can be expressed in terms of the hypergeometric function as given in Eq. (26). Equation (40) follows from Eq. (29) in [25] and allows calculating the mathematical limits \(\eta \rightarrow 0\) and \(\eta \rightarrow 1\).

The frequency distribution \(p(\omega )\) is defined as the Fourier transform of the free induction decay:

$$\begin{aligned} p(\omega ) = \frac{1}{2\pi }\int \limits _{-\infty }^{+\infty } \mathrm {d}t M(t) \mathrm {e}^{-\mathrm {i}\omega t}\,. \end{aligned}$$
(41)

Since \(M(0) = 1\) [see Eq. (3)], \(p(\omega )\) is normalized to:

$$\begin{aligned} \int \limits _{-\infty }^{+\infty } \mathrm {d}\omega \, p(\omega ) = 1\,. \end{aligned}$$
(42)

The frequency distribution in the static dephasing limit for the SVM was found as [18]:

$$\begin{aligned} p(\omega ) = \frac{\eta }{1-\eta } \frac{\delta \omega }{\pi \omega ^2} \left\{ \begin{array}{ll} \sqrt{1-\left[ \frac{\omega }{\delta \omega }\right] ^2} &{\quad} \text {for } -\delta \omega \le \omega \le -\eta \delta \omega \; \text {or} \; +\eta \delta \omega \le \omega \le +\delta \omega \\ \\ \sqrt{1-\left[ \frac{\omega }{\delta \omega }\right] ^2} - \sqrt{1-\left[ \frac{\omega }{\eta \, \delta \omega }\right] ^2} &{\quad} \text {for } -\eta \delta \omega<\omega < +\eta \delta \omega \\ \\ 0 &{\quad} \text {elsewhere} \,. \end{array} \right. \end{aligned}$$
(43)

Evidently, \(p(\omega )\) exhibits two prominent peaks at \(\omega = \pm \eta \delta \omega \). The frequency distribution from Eq. (43) possesses a local minimum at \(\omega = 0\), whereas \(p(\omega )\) in the RVM peaks globally at \(\omega = 0\) (as shown in the blue lines in Fig. 6). The mean relaxation time approach to determine the relaxation rate \(R_2^{*}\) [see Eq. (4)], however, translates into the Fourier domain as follows [91]:

$$\begin{aligned} R_2^{*} = \frac{1}{\pi p(0)} = \frac{2\eta \delta \omega }{1+\eta }\,. \end{aligned}$$
(44)

Thus, it follows that the relaxation rate for the SVM in the static dephasing regime is twice as large as that for the RVM, while both frequency distributions exhibit a similar lineshape, i.e. both distributions have the same width:

$$\begin{aligned} \langle \omega ^2 \rangle = \int \limits _{-\infty }^{+\infty } \mathrm {d}\omega p(\omega ) \omega ^2 = \frac{\eta }{2}\delta \omega ^2\,. \end{aligned}$$
(45)

In the limit of static dephasing, Yablonskiy and Haacke could connect the free induction decay of N randomly distributed capillaries, \(M_{0,\text {RVM}}\), with the free induction decay around one single blood vessel, \(M_{0,\text {SVM}}\), as in [10]:

$$\begin{aligned} M_{0,\text {RVM}}(t) = \left[ M_{0,\text {SVM}}\left( \frac{\eta }{N},t\right) \right] ^N\,, \end{aligned}$$
(46)

where \(M_{0,\text {SVM}}\) is given in Eq. (39). In the statistical limit \(N\rightarrow \infty \), the gradient echo signal in the RVM is then given in Eq. (10).

Exact solution of the Bloch-Torrey equation in the SVM

An analytical solution of the local magnetization and the free induction decay in the SVM can be obtained with an eigenfunction expansion of the Bloch-Torrey equation [22, 24].

$$\begin{aligned} M_{\text {SVM}}(t) = \sum \limits _{m=0}^{\infty }\sum \limits _{n=1}^{\infty }d_{nm}\mathrm {e}^{-\lambda _{nm}^2\frac{Dt}{R^2}}\,. \end{aligned}$$
(47)

The eigenvalues \(\lambda _{nm}\) in the SVM are, in general, complex and obey the following condition:

$$\begin{aligned} Y^{\prime }_{k_m}(\lambda _{nm}) J^{\prime }_{k_m}\left( \frac{\lambda _{nm}}{\sqrt{\eta }}\right) = J^{\prime }_{k_m}(\lambda _{nm}) Y^{\prime }_{k_m}\left( \frac{\lambda _{nm}}{\sqrt{\eta }}\right) \,. \end{aligned}$$
(48)

Here \(J_{k_m}\) and \(Y_{k_m}\) are Bessel functions of the first and second kind and \(J^{\prime }_{k_m}\) and \(Y^{\prime }_{k_m}\) are derivations of Bessel functions, respectively. The index \(k_m\) relates to the interplay between diffusion and susceptibility effects, since:

$$\begin{aligned} k_m^2 = a_{2m} \left( \frac{\mathrm {i}}{2}\delta \omega \frac{R^2}{D}\right) \,, \end{aligned}$$
(49)

where \(a_{2m}\) denotes the characteristic values of the Mathieu functions [92]. The expansion coefficients \(d_{nm}\) are given as:

$$\begin{aligned} d_{nm}=\frac{8\eta }{1-\eta }\left[ A_0^{(2m)} \right] ^2\frac{\left[ J^{\prime }_{k_m}(\lambda _{nm}) s^{\prime }_{1,k_m}\left( \frac{\lambda _{nm}}{\sqrt{\eta }}\right) -J^{\prime }_{k_m}\left( \frac{\lambda _{nm}}{\sqrt{\eta }}\right) s^{\prime }_{1,k_m}\left( \lambda _{nm}\right) \right] ^2}{\left[ J^{\prime }_{k_m}(\lambda _{nm})\right] ^2 \left[ \lambda _{nm}^2-\eta k_m^2 \right] -\left[ J^{\prime }_{k_m}\left( \frac{\lambda _{nm}}{\sqrt{\eta }}\right) \right] ^2 \left[ \lambda _{nm}^2-k_m^2 \right] }\,, \end{aligned}$$
(50)

with \(s^{\prime }_{1,k_m}\) denoting the first derivative of the Lommel function \(s_{1,k_m}\) (see also [93]). The symbol \(A_0^{(2m)}\) represents the first Fourier coefficient of the even \(\pi \)-periodic Mathieu functions \(\text {ce}_{2m}\) and also depends on the parameter \(\delta \omega R^2/D\). This expansion uses angular eigenvalues \(\kappa _m\) and associated eigenvectors with an index m, as well as radial eigenvalues \(\lambda_{nm}\) and associated eigenvectors with indices m and n. The indices n and m are therefore analogues of the radial and angular quantum numbers in quantum mechanics. The total magnetization is then given as sum over all eigenfunctions and corresponding eigenvalues as given in Eq. (47). One could argue against such an expansion since the Bloch-Torrey operator, \(D\Delta -\mathrm {i}\omega \), is non-hermitian. However, an operator does not necessarily need to be hermitian to allow for an eigenfunction expansion as e.g. exemplified in [94, 95]. Furthermore, Herberthson et al. provided proof that the eigenfunctions of the Bloch-Torrey operator for a linear field gradient are self-dual [96]. This proof can easily be generalized for arbitrary magnetic field inhomogeneities for reflecting boundary conditions:

Lemma 1

Suppose u and v are eigenfunctions of the operator \(\Delta +z(\mathbf {r})\) with the eigenvalues \(\upmu \) and \(\lambda \), respectively, which fulfill reflecting boundary conditions at the surface \(\partial \Omega \) of the volume \(\Omega \). The function \(z(\mathbf {r})\in \mathbb {C}\) is an arbitrary complex function. Then, the eigenfunctions are self-dual, e.g. if \(\upmu \ne \lambda \), then\(\int _\Omega \mathrm {d}\Omega \, uv = 0\).

Proof

Suppose \(\upmu \ne \lambda \).

$$\begin{aligned} \upmu \int _{\Omega } \mathrm {d}\Omega \,vu&= \int _{\Omega } \mathrm {d}\Omega \,v\left[ \Delta u +z(\mathbf {r}) u \right] \end{aligned}$$
(51)
$$\begin{aligned}&= \int _{\Omega } \mathrm {d}\Omega \,u \left[ \Delta v\right] - \underbrace{\int _{\partial \Omega } \hat{n} \mathrm {d}s [v\nabla u-u\nabla v]}_{=0}+\int _{\Omega } \mathrm {d}\Omega \,z(\mathbf {r})\, u v\,, \end{aligned}$$
(52)

where the second term vanishes due to the reflecting boundary conditions and Green’s identity was applied. Hereby, \(\hat{n}\) denotes a unity vector that is orthogonal to the surface \(\partial \Omega \) of the volume \(\Omega \), \(\mathrm {d}\Omega \) symbolizes the volume integration and \(\mathrm {d}s\) symbolizes the integration over the surface \(\partial \Omega \). Finally, one obtains:

$$\begin{aligned} \upmu \int _{\Omega } \mathrm {d}\Omega \,uv = \int _{\Omega } \mathrm {d}\Omega \,u [\Delta +z(\mathbf {r})]v = \lambda \int _{\Omega } \mathrm {d}\Omega \,uv\,. \end{aligned}$$
(53)

Thus, the statement follows. \(\square \)

In the special case \(z(\mathbf {r}) \propto -\mathrm {i}\omega (\mathbf {r})\), where the function z takes purely imaginary values, the eigenfunctions of the Bloch-Torrey operator are self-dual. This agrees with Eq. (4) in [92] and Eqs. (9) and (12) in [24]. In addition, an actual eigenfunction expansion for the Bloch-Torrey equation has been successfully developed in several publications, e.g. [97,98,99,100]. A more detailed account is given in [101].

Fig. 6
figure 6

Frequency distributions for \(\eta = 0.05\) and \(\delta \omega = 1000/\text {s}\), which correspond to \(B_0=9.4\, \text {T}\) and \(D=1\,\upmu \text {m}^2/\text {ms}\) for different capillary radii. The solid lines present the strong collision approximation (SCA) of the randomly positioned capillary model obtained from Eq. (82), and exact solutions in the single-vessel approximation are shown as dashed lines [see Eq. (47)]. In the static dephasing limit, both distributions exhibit the same width [see Eq. (45)]. The frequency distribution in the SVM, however, possesses a local minimum at \(\omega = 0\) for large capillary radii, whereas the RVM contains a global maximum at \(\omega = 0\). This difference gives different estimations of the relaxation rate \(R_2^{*}\), as shown in Eq. (44)

According to Eq. (4), the relaxation rate in this model can be found as:

$$\begin{aligned} R_2^{*} = \frac{D}{R^2}\left[ \sum \limits _{m=0}^{\infty }\sum \limits _{n=1}^{\infty } \frac{d_{nm}}{\lambda _{nm}^2}\right] ^{-1}\,. \end{aligned}$$
(54)

Strong collision approximation (SCA)

In the static dephasing limit, the free induction decay is given in Eq. (39); its Laplace transform then follows as:

$$\begin{aligned} \hat{M}_{0,\text {SVM}}(s)=\frac{1}{1-\eta }\frac{1}{s} \left[ \sqrt{1+\frac{\eta ^2\delta \omega ^2}{s^2}}-\eta \sqrt{1+\frac{\delta \omega ^2}{s^2}}\right] \,, \end{aligned}$$
(55)

where the \(\hat{}\) symbol denotes the Laplace transform of a quantity as defined in Eq. (72) in “Appendix C”. We then find the SCA relaxation rate as [25]:

$$\begin{aligned} R_2^{*}=\frac{\sqrt{1+\left[ \eta \tau _{\text {SVM}}\delta \omega \right] ^2}+\eta \sqrt{1+\left[ \tau _{\text {SVM}}\delta \omega \right] ^2}}{\tau _{\text {SVM}}[1+\eta ]}-\frac{1}{\tau _{\text {SVM}}}\,. \end{aligned}$$
(56)

In the SVM, the correlation time \(\tau \), as given in Eq. (73) in “Appendix C”, can be obtained as [18]:

$$\begin{aligned} \tau _{\text {SVM}} =\frac{R^2}{4D}\frac{\mathrm {ln}(\eta )}{\eta -1}\, \end{aligned}$$
(57)

and the spin echo relaxation rate is determined by [19]:

$$\begin{aligned} R_2 = \frac{R_2^{*}}{1+\tau R_2^{*}}\,, \end{aligned}$$
(58)

with the gradient echo relaxation rate \(R_2^{*}\) from Eq. (56) and the correlation time \(\tau \) from Eq. (57). Thus, the spin echo relaxation rate in the SCA reads:

$$\begin{aligned} R_2 = \tau _{\text {SVM}}^{-1} \left[ 1-\frac{1+\eta }{\sqrt{1+[\eta \tau _{\text {SVM}}\delta \omega ]^2}+\eta \sqrt{1+[\tau _{\text {SVM}}\delta \omega ]^2}} \right] \,. \end{aligned}$$
(59)

Appendix B: Weak field approximation

In analogy to Jensen et al. [15], the correlation function for a randomized distribution of permeable capillaries is given as [18, 38, 106]:

$$\begin{aligned} K(t) = \frac{1}{V}\int _V \mathrm {d}^2 \mathbf {r} \int _V \mathrm {d}^2 \mathbf {s}\, \omega (\mathbf {r}) \frac{1}{4\pi D t} \mathrm {e}^{-\frac{|\mathbf {r}-\mathbf {s}|^2}{4Dt}} \omega (\mathbf {s})\,, \end{aligned}$$
(60)

where the free diffusion of spin-bearing particles was assumed, which is a good approximation for small capillary blood volume fractions \(\eta \). The correlation function at the initial time point \(t = 0\) takes the value: \(K(0) = \langle \omega ^2(\mathbf {r}) \rangle \). Using the local Larmor frequency given in Eq. (7), we can simplify K(t) to:

$$\begin{aligned} K(t) = \frac{\eta \delta \omega ^2 R^2}{4\pi ^2 D t }\int _{|\mathbf {r}|>R} \mathrm {d}^2 \mathbf {r}\int _{|\mathbf {s}|>R} \mathrm {d}^2 \mathbf {s} \,\mathrm {e}^{-\frac{|\mathbf {r}-\mathbf {s}|^2}{4Dt}} \frac{2[\hat{x}\mathbf {r}]^2-r^2}{r^4}\frac{2[\hat{x}\mathbf {s}]^2-s^2}{s^4}\,, \end{aligned}$$
(61)

where \(\hat{x}\) is the unit vector in the spatial x direction and a random distribution of capillaries in a plane is assumed. This approximation, however, allows for an overlap of capillary positions. Yet, for low capillary blood volume fractions \(\eta \), the occurrences of capillary overlaps are very scarce and, thus, can be neglected. With the representation of the diffusion propagator in Fourier space

$$\begin{aligned} \mathrm {e}^{-\frac{|\mathbf {r}-\mathbf {s}|^2}{4Dt}} = \frac{Dt}{\pi } \int \mathrm {d}^2 \mathbf {k}\, \mathrm {e}^{\mathrm {i}\mathbf {k}[\mathbf {r}-\mathbf {s}]}\mathrm {e}^{-Dtk^2}, \end{aligned}$$
(62)

we find

$$\begin{aligned} K(t)&= \frac{\eta \delta \omega ^2 R^2}{4\pi ^3} \int \mathrm {d}^2 \mathbf {k} \mathrm {e}^{-Dtk^2} \left| \int _{r>R} \mathrm {d}^2 \mathbf {r} \frac{2[\hat{x}\mathbf {r}]^2-r^2}{r^4} \mathrm {e}^{\mathrm {i}\mathbf {k}\mathbf {r} } \right| ^2\end{aligned}$$
(63)
$$\begin{aligned}& \qquad= \frac{\eta \delta \omega ^2 R^2}{4\pi ^3} \int \mathrm {d}^2 \mathbf {k} \mathrm {e}^{-Dtk^2} \left| \int _{R}^{\infty } \mathrm {d}r r \int _{0}^{2\pi } \mathrm {d}\phi \frac{\cos (2\phi )}{r^2} \mathrm {e}^{\mathrm {i}k r \cos (\phi -\vartheta )} \right| ^2\,, \end{aligned}$$
(64)

where \(\phi \) is the angle between the position vector \(\mathbf {r}\) and the transverse component of the magnetic field (see Fig. 1) and \(\vartheta \) denotes the angle between the reciprocal vector \(\mathbf {k}\) and the magnetic field. Evaluating the angular part \(\phi \) of the integral, we find:

$$\begin{aligned} K(t)&= \frac{\eta \delta \omega ^2 R^2}{\pi } \int \mathrm {d}^2\mathbf {k} \cos ^2(2\vartheta ) \mathrm {e}^{-Dtk^2} \left[ \int _{R}^{\infty } \frac{\mathrm {d}r}{r} J_2(kr)\right] ^2\end{aligned}$$
(65)
$$\begin{aligned}&=\frac{\eta \delta \omega ^2}{\pi } \int _{0}^{2\pi } \mathrm {d}\vartheta \cos ^2(2\vartheta ) \int _{0}^{\infty } \frac{\mathrm {d}k}{k}\mathrm {e}^{-Dtk^2} [J_1(kR)]^2\,, \end{aligned}$$
(66)

where \(J_1\) and \(J_2\) are Bessel functions of the first kind. By evaluating the integrals over \(\vartheta \) and k, the correlation function follows as:

$$\begin{aligned} K(t)= \frac{\eta \delta \omega ^2}{2} \left[ 1-\mathrm {e}^{-\frac{R^2}{2Dt}}\left[ I_0\left( \frac{R^2}{2Dt}\right) +I_1\left( \frac{R^2}{2Dt}\right) \right] \right] \,, \end{aligned}$$
(67)

where \(I_0\) and \(I_1\) are modified Bessel functions of the first kind. The prefactor \(\eta \delta \omega ^2/2 = \langle \omega ^2 \rangle \) equals the width of the frequency distribution as demonstrated below [see Eq. (45)]. For very long or short times, the correlation function can be approximated as:

$$\begin{aligned} K(t) \approx {\left\{ \begin{array}{ll} \frac{\eta \delta \omega ^2}{2}\,& \text {for}\, \frac{R^2}{2D}\gg t \\ \\ \frac{\eta \delta \omega ^2R^2}{8Dt} \,& \text {for}\, \frac{R^2}{2D}\ll t\,. \end{array}\right. } \end{aligned}$$
(68)

According to [15], the weak field approximation (WFA) is valid for

$$\begin{aligned} R_2^{*} < \frac{D}{R^2}\,. \end{aligned}$$
(69)

Since \(R_2^{*} \approx \frac{\eta }{2}\tau \delta \omega ^2\) in the motional narrowing regime, the WFA is valid for \(\eta \frac{R^4}{D^2}\delta \omega ^2 < 3.72\). Non-Gaussian spin behaviour was analyzed in [107].

In Fig. 7, the calculated correlation function is compared with the correlation function of Sukstanskii and Yablonskiy as obtained in the RVM for impermeable vessel walls [17, 38]:

$$\begin{aligned} K(t) = \frac{16}{\pi ^2}\eta \delta \omega ^2\int \limits _{0}^{\infty } \frac{\mathrm {d}x}{x^5} \frac{\exp \left( -\frac{Dt}{R^2}x^2\right) }{\left[ {J_2^{\prime }}(x)\right] ^2+\left[ {Y_2^{\prime }}(x)\right] ^2}\,, \end{aligned}$$
(70)

where \(J_2^{\prime }\) is the first derivative of the Bessel functions of the first kind with index 2 and \(Y_2^{\prime }\) is the first derivative of the Bessel functions of the second kind with index 2.

Fig. 7
figure 7

Comparison of the correlation function in the RVM for permeable vessel walls, given in Eq. (67), with the correlation function in the RVM for impermeable vessels (yellow line), given in Eq. (70) and the correlation function in the SVM (see Ref. [106], red line). All functions coincide for short times. The blue dotted line shows the monoexponential approximation of the correlation function with the correlation time given in Eq. (76)

As shown in Fig. 7, the correlation function in the RVM for impermeable vessels is similar to the correlation function in the RVM for permeable vessels. Especially, both curves exhibit the same asymptotic behaviour for long times that is shown in Eq. (68).

Random walk simulations were performed to validate the correlation functions. Therefore, 100,000 random walk trajectories were simulated in a volume with 17,500 randomly distributed cylindrical vessels with capillary volume fraction \(\eta =0.05\). The obtained correlation function is compared in Fig. 7 with the correlation functions in the SVM and RVM for permeable and impermeable vessels.

Appendix C: strong collision approximation (SCA) in the random vessel distribution model (RVM)

In the SCA, the diffusion operator in the Bloch-Torrey equation is replaced by a simpler Markov operator. In a Markov process, the diffusion transition rate between different positions is only dependent on the equilibrium probability of the final state. One of the main results of the SCA condenses in the connection between the Laplace transform of the free induction decay with the Laplace transform of the static dephasing free induction decay and the correlation time \(\tau \) [see Eq. (20) and [28]]:

$$\begin{aligned} \hat{M}(s) = \frac{\hat{M}_0(s+\tau ^{-1})}{1-\tau ^{-1}\hat{M}_0(s+\tau ^{-1})}\,, \end{aligned}$$
(71)

where the Laplace transform is defined as:

$$\begin{aligned} \hat{M}(s) = \int \limits _{0}^{\infty } \mathrm {d}t \mathrm {e}^{-st} M(t)\,. \end{aligned}$$
(72)

Lerch’s uniqueness theorem guarantees the uniqueness of the inverse Laplace transform of \(\hat{M}(s)\), even if the function M(t) is only piecewise continuous [102]. With the strong collision assumption, the correlation function exhibits monoexponential decay with the correlation time \(\tau \) of the form [107]:

$$\begin{aligned} K(t) = K(0) \mathrm {e}^{-\frac{t}{\tau }}\,. \end{aligned}$$
(73)

The correlation time \(\tau \) can then be obtained with the least squares method by minimizing the functional difference between the correlation function K(t) with its monoexponential approximation [see Eq. (19)]. For permeable vessel boundaries, we obtained with the correlation function K(t) in Eq. (67) the correlation time \(\tau _\text {per}\) as a solution of the equation:

$$\begin{aligned} \frac{3}{4}\sqrt{\pi } = G_{1,3}^{2,1}\left( \frac{R^2}{D}\tau _{\text {per}}^{-1}\left |\frac{1}{2}; 0,2,0 \right) + G_{1,3}^{2,1}\left( \frac{R^2}{D}\tau _{\text {per}}^{-1}\right|\frac{1}{2}; 1,2,-1 \right) \,, \end{aligned}$$
(74)

where G denotes Meijers G-function. The correlation time \(\tau _{\text {per}}\) can then obtained as

$$\begin{aligned} \tau _{\text {per}} \approx 0.5374\,\frac{R^2}{D}\,. \end{aligned}$$
(75)

For impermeable vessels, the correlation time \(\tau \) is also defined by minimizing the expression given in Eq. (19):

$$\begin{aligned} \tau \approx 0.6902\frac{R^2}{D}\,, \end{aligned}$$
(76)

where K(t) is given in Eq. (70). Since impermeable vessel walls provide a more realistic tissue model than permeable vessels walls, the correlation time of impermeable vessels is used for the SCA in the RVM.

Details about the derivation of Eq. (71) are presented in appendix A in Ref. [18], or in [64, 103, 104]. For the static dephasing magnetization in the RVM given in Eq. (28), the Laplace transform follows as:

$$\begin{aligned} \hat{M}_{0,\text {RVM}}(s) = \sqrt{\frac{\pi }{\eta }}\frac{\mathrm {e}^{\frac{s^2}{\eta \delta \omega ^2}}}{\delta \omega }\mathrm {erf}\left( \frac{s}{\sqrt{\eta }\delta \omega },\frac{s}{\sqrt{\eta }\delta \omega }+\frac{3\sqrt{\eta }}{4}\right) +\frac{\mathrm {e}^{-\frac{3s}{2\delta \omega }-\frac{\eta }{2}}}{\eta \delta \omega +s}\,, \end{aligned}$$
(77)

where \(\mathrm {erf}(z_0,z_1) = \mathrm {erf}(z_1)- \mathrm {erf}(z_0)\). When inserting the Laplace transform of the RVM static dephasing free induction decay into Eq. (71), one finds:

$$\begin{aligned} \hat{M}_{\text {RVM}}(s) = \tau \frac{\frac{1}{\tau \delta \omega }\sqrt{\frac{\pi }{\eta }}\mathrm {e}^{\frac{1}{\eta }\left[ \frac{\tau s+1}{\tau \delta \omega }\right] ^2}\mathrm {erf}\left( \frac{\tau s+1}{\sqrt{\eta }\tau \delta \omega },\frac{\tau s+1}{\sqrt{\eta }\tau \delta \omega }+\frac{3\sqrt{\eta }}{4}\right) +\frac{\mathrm {e}^{\frac{3\tau s+3-\eta \tau \delta \omega }{2\tau \delta \omega }}}{\eta \tau \delta \omega + \tau s+1}}{1- \frac{1}{\tau \delta \omega }\sqrt{\frac{\pi }{\eta }}\mathrm {e}^{\frac{1}{\eta }\left[ \frac{\tau s+1}{\tau \delta \omega }\right] ^2}\mathrm {erf}\left( \frac{\tau s+1}{\sqrt{\eta }\tau \delta \omega },\frac{\tau s+1}{\sqrt{\eta }\tau \delta \omega }+\frac{3\sqrt{\eta }}{4}\right) -\frac{\mathrm {e}^{\frac{3\tau s+3-\eta \tau \delta \omega }{2\tau \delta \omega }}}{\eta \tau \delta \omega +\tau s+1}}\,. \end{aligned}$$
(78)

The monoexponential relaxation rate \(R_2^{*}\) is defined in Eq. (4):

$$\begin{aligned} R_2^{*}&=\left[ \hat{M}(0)\right] ^{-1}\end{aligned}$$
(79)
$$\begin{aligned}&= \frac{1}{\hat{M}_0(\tau ^{-1})} - \tau ^{-1}\,. \end{aligned}$$
(80)

Inserting the definition of the Laplace transform given in Eq. (72), one finds Eq. (20). Finally, the gradient echo relaxation rate can be obtained as:

$$\begin{aligned} R_2^*=\frac{\delta \omega }{\mathrm {e}^{\frac{1}{\eta [\tau \delta \omega ]^2}}\sqrt{\frac{\pi }{\eta }}\mathrm {erf}\left( \frac{1}{\sqrt{\eta }\tau \delta \omega },\frac{1}{\sqrt{\eta }\tau \delta \omega }+\frac{3\sqrt{\eta }}{4}\right) +\frac{\tau \delta \omega }{\eta \tau \delta \omega +1}\mathrm {e}^{-\frac{3}{2\tau \delta \omega }-\frac{\eta }{2}}}-\frac{1}{\tau }\,. \end{aligned}$$
(81)

A Taylor expansion with respect to the capillary blood volume fraction \(\eta \) leads to Eq. (30). Both gradient echo relaxation rate \(R_2^{*}\) and spin echo relaxation rate \(R_2\) are shown in Figs. 8 and 9 for different magnetic field strengths \(B_0\), for a blood volume fraction of \(\eta = 0.05\), a dipole field strength of \(\delta \omega \approx 100 \frac{1}{\text {s} \text {T}} B_0\) in accordance with Eq. (8) and a diffusion coefficient of \(D=1\,\upmu \text {m}^2 \text {ms}^{-1}\) in accordance with [41]. A change of the blood volume fraction would linearly scale the ordinates in Figs. 8 and 9, while the abscissas show a dependence on the product \(\delta \omega R^2/D\) [105].

Fig. 8
figure 8

Gradient echo relaxation rate \(R_2^{*}\) for different magnetic field strength \(B_0\). The static dephasing regime is valid for large vessel radii and the weak field approximation in the opposite limit for small vessel radii. The strong collision approximation (SCA) agrees very well with random walk simulations for arbitrary vessel radii

Fig. 9
figure 9

Spin echo relaxation rate \(R_2\) for different magnetic field strengths \(B_0\). The slow diffusion approximation is valid for large vessel radii and its accuracy increases with increasing magnetic field strengths. The weak field approximation is valid in the opposite limit for small vessel radii and small magnetic field strengths. The strong collision approximation (SCA) can only qualitatively describe the shape of the \(R_2\) simulation data points

With the Laplace transform of the free induction decay, \(\hat{M}_{\text {RVM}}(s)\), the frequency distribution follows as:

$$\begin{aligned} p(\omega ) = \frac{1}{\pi } \mathrm{Re} \left( \hat{M}(\mathrm {i}\omega )\right) \,. \end{aligned}$$
(82)

In Fig. 6, the frequency distribution is shown for different capillary radii; evidently, with increasing capillary radius, the frequency distribution broadens.

So far, the gradient echo signal was analyzed for a fixed angle \(\theta \) between the main magnetic field and vessel orientation. An averaging over all possible orientations (index RO) is therefore a natural step to account for the intrinsic inhomogeneities of blood vessel networks. In this limit, Yablonskiy and Haacke obtained an expression for the static dephasing magnetization given in Eq. (33). Using a similar procedure as above, we can obtain the Laplace transform

$$\hat{M}_{0,\text {RO}}(s) = \frac{\mathrm {e}^{- \frac{3s}{2\overline{\delta \omega }}-\frac{\eta }{2}}}{s+\eta \overline{\delta \omega }}+\sqrt{\frac{5\pi }{6\eta }}\frac{ \mathrm{e}^{\frac{5s^2}{6\eta \overline{\delta \omega }^2 }}}{\overline{\delta \omega }}\text {erf}\left(\frac{ s}{\overline{\delta \omega }}\sqrt{\frac{5}{6\eta }},\frac{s}{\overline{\delta \omega}}\sqrt{\frac{5}{6\eta }}+\frac{9}{2}\sqrt{\frac{\eta }{30}}\right)$$
(83)

and, eventually, the SCA relaxation rate \(\overline{R_2^{*}}\) as:

$$\begin{aligned} \tau \overline{R_2^*}= \left[ \sqrt{\frac{5\pi }{6\eta }}\frac{\mathrm {e}^{\frac{5}{6 \eta [\tau \overline{\delta \omega }]^2}}}{\tau \overline{\delta \omega }} \text {erf}\left( \frac{1}{\tau \overline{\delta \omega }}\sqrt{\frac{5}{6\eta }},\frac{1}{\tau \overline{\delta \omega }}\sqrt{\frac{5}{6\eta }}+\frac{9}{2}\sqrt{\frac{\eta }{30}}\right) +\frac{\mathrm {e}^{-\frac{3}{2\tau \overline{\delta \omega }}-\frac{\eta }{2}}}{1+\eta \tau \overline{\delta \omega }}\right] ^{-1}-1. \end{aligned}$$
(84)

A Taylor expansion in order of \(\eta \) again leads to Eq. (34).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Buschle, L.R., Ziener, C.H., Zhang, K. et al. Vessel radius mapping in an extended model of transverse relaxation. Magn Reson Mater Phy 31, 531–551 (2018). https://doi.org/10.1007/s10334-018-0677-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10334-018-0677-9

Keywords

Navigation