random-walk technique for simulating nmr measurements and ... · 90 along the x^0 axis, followed by...

14
Random-walk technique for simulating NMR measurements and 2D NMR maps of porous media with relaxing and permeable boundaries Emmanuel Toumelin a,1 , Carlos Torres-Verdı ´n a, * , Boqin Sun b , Keh-Jim Dunn b a Department of Petroleum and Geosystems Engineering, The University of Texas, Austin, TX 78712, USA b Chevron Energy Technology Company, San Ramon, CA 94583, USA Received 18 July 2006; revised 23 May 2007 Available online 17 June 2007 Abstract We revisit random-walk methods to simulate the NMR response of fluids in porous media. Simulations reproduce the effects of dif- fusion within external inhomogeneous background magnetic fields, imperfect and finite-duration B 1 pulses, T 1 /T 2 contrasts, and relaxing or permeable boundaries. The simulation approach consolidates existing NMR numerical methods used in biology and engineering into a single formulation that expands on the magnetic-dipole equivalent of spin packets. When fluids exhibit low T 1 /T 2 contrasts and when CPMG pulse sequences are used to acquire NMR measurements, we verify that classical NMR numerical models that neglect T 1 effects accurately reproduce surface magnetization decays of saturated granular porous media regardless of the diffusion/relaxation regime. Currently, analytical expressions exist only for the case of arbitrary pore shapes within the fast-diffusion limit. However, when fluids include several components or when magnetic fields are strongly inhomogeneous, we show that simulations results obtained using the complete set of Bloch’s equations differ substantially from those of classical NMR models. In addition, our random-walk formulation accurately reproduces magnetization echoes stemming from coherent-pathway calculations. We show that the random-walk approach is especially suited to generate parametric multi-dimensional T 1 /T 2 /D NMR maps to improve the characterization of pore structures and saturating fluids. Ó 2007 Elsevier Inc. All rights reserved. Keywords: Random walks; NMR relaxation; Porous media; Fluid typing; 2D NMR 1. Introduction The application of NMR measurements for pore size characterization of porous and permeable media is moti- vated by two relationships between measurable quantities and the pore surface-to-volume ratio S/V of porous media saturated with a single fluid phase. First, longitudinal and transversal relaxation times of a fluid (T 1 and T 2 , respec- tively) are related to the product q(S/V), where q is the relaxivity of the pore surface. This relationship remains valid in the fast-diffusion limit [1] when paramagnetic cen- ters are present at the pore wall surface. Second, the short- time asymptotic behavior of the effective fluid diffusivity D is directly related to S/V [2,3]. Practical limitations exist when applying these two relationships to the interpretation of NMR measurements of porous media. Specifically, it is normally assumed that (1) pores are well defined and iso- lated, and (2) inversion techniques can reliably and accu- rately estimate the distribution of surface-to-volume ratios. Experimental evidence shows that pores spanning different length scales can be connected and, therefore, that the estimated values of S/V could be interpreted in an aver- age sense in the presence of low surface relaxivity, such as in certain carbonate rocks [4]. In addition, effective pore values of S/V sensed by protons in a fluid phase depend 1090-7807/$ - see front matter Ó 2007 Elsevier Inc. All rights reserved. doi:10.1016/j.jmr.2007.05.024 * Corresponding author. Fax: +1 512 471 4900. E-mail address: [email protected] (C. Torres-Verdı ´n). 1 Present address: Chevron North America Exploration and Production, Houston, TX 77099, USA. www.elsevier.com/locate/jmr Journal of Magnetic Resonance 188 (2007) 83–96

Upload: others

Post on 04-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

www.elsevier.com/locate/jmr

Journal of Magnetic Resonance 188 (2007) 83–96

Random-walk technique for simulating NMR measurementsand 2D NMR maps of porous media with relaxing and

permeable boundaries

Emmanuel Toumelin a,1, Carlos Torres-Verdın a,*, Boqin Sun b, Keh-Jim Dunn b

a Department of Petroleum and Geosystems Engineering, The University of Texas, Austin, TX 78712, USAb Chevron Energy Technology Company, San Ramon, CA 94583, USA

Received 18 July 2006; revised 23 May 2007Available online 17 June 2007

Abstract

We revisit random-walk methods to simulate the NMR response of fluids in porous media. Simulations reproduce the effects of dif-fusion within external inhomogeneous background magnetic fields, imperfect and finite-duration B1 pulses, T1/T2 contrasts, and relaxingor permeable boundaries. The simulation approach consolidates existing NMR numerical methods used in biology and engineering intoa single formulation that expands on the magnetic-dipole equivalent of spin packets. When fluids exhibit low T1/T2 contrasts and whenCPMG pulse sequences are used to acquire NMR measurements, we verify that classical NMR numerical models that neglect T1 effectsaccurately reproduce surface magnetization decays of saturated granular porous media regardless of the diffusion/relaxation regime.Currently, analytical expressions exist only for the case of arbitrary pore shapes within the fast-diffusion limit. However, when fluidsinclude several components or when magnetic fields are strongly inhomogeneous, we show that simulations results obtained using thecomplete set of Bloch’s equations differ substantially from those of classical NMR models. In addition, our random-walk formulationaccurately reproduces magnetization echoes stemming from coherent-pathway calculations. We show that the random-walk approach isespecially suited to generate parametric multi-dimensional T1/T2/D NMR maps to improve the characterization of pore structures andsaturating fluids.� 2007 Elsevier Inc. All rights reserved.

Keywords: Random walks; NMR relaxation; Porous media; Fluid typing; 2D NMR

1. Introduction

The application of NMR measurements for pore sizecharacterization of porous and permeable media is moti-vated by two relationships between measurable quantitiesand the pore surface-to-volume ratio S/V of porous mediasaturated with a single fluid phase. First, longitudinal andtransversal relaxation times of a fluid (T1 and T2, respec-tively) are related to the product q(S/V), where q is therelaxivity of the pore surface. This relationship remains

1090-7807/$ - see front matter � 2007 Elsevier Inc. All rights reserved.

doi:10.1016/j.jmr.2007.05.024

* Corresponding author. Fax: +1 512 471 4900.E-mail address: [email protected] (C. Torres-Verdın).

1 Present address: Chevron North America Exploration and Production,Houston, TX 77099, USA.

valid in the fast-diffusion limit [1] when paramagnetic cen-ters are present at the pore wall surface. Second, the short-time asymptotic behavior of the effective fluid diffusivity D

is directly related to S/V [2,3]. Practical limitations existwhen applying these two relationships to the interpretationof NMR measurements of porous media. Specifically, it isnormally assumed that (1) pores are well defined and iso-lated, and (2) inversion techniques can reliably and accu-rately estimate the distribution of surface-to-volumeratios. Experimental evidence shows that pores spanningdifferent length scales can be connected and, therefore, thatthe estimated values of S/V could be interpreted in an aver-age sense in the presence of low surface relaxivity, such asin certain carbonate rocks [4]. In addition, effective porevalues of S/V sensed by protons in a fluid phase depend

Page 2: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

84 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

on the relative saturation of that phase within the pore:when several fluid phases are present in the same porespace, such as in the case of mixed-wettabilities, the actualand effective values of S/V for that pore may differ substan-tially. Models used in biological applications to study theinfluence of membrane permeability on NMR signals alsoover-simplify the pore structure of biological systemsbecause of the lack of efficient numerical simulation tech-niques [5–8].

Recently, special acquisition techniques have been devel-oped to improve the accuracy of NMR measurements toassess fluid types [9–11] using (a) suites of magnetic pulsesequences of different magnitudes, orientations, and char-acteristic times, e.g., CPMG-like sequences with differentecho times and wait times, and (b) multi-dimensional inver-sion of the measured magnetization decays in the form ofintensity maps in T1, T2, and D domains. Despite theseefforts, ambiguities remain when the saturating fluids exhi-bit similar characteristics, such as light oil and water inwater-wet and mixed-wet rocks, because the inversion pro-cess is sensitive to fluid diffusivity contrasts, restricted dif-fusion, surface relaxation, and wettability of the porousmedium.

Modern NMR applications require simulation andinterpretation methods that can incorporate the complexityof real porous media and multiphase fluid saturation at thepore scale. Accurate simulation methods have been pro-posed based on coherent pathways for bulk fluids [12–15], or based on analytical diffusion propagators for thecase of periodic porous microstructures [16]. However,the disordered nature of most biological and mineral por-ous media demands more sophisticated approaches fornumerical simulation.

The objective of this paper is to develop an accurate anefficient algorithm for the simulation of NMR measure-ments of saturated microporous media. Specifically, thealgorithm is designed to account for known distributionsof (a) pore boundaries and surface characteristics (i.e.,geometry and surface relaxation q and/or diffusion perme-ability P), (b) static and susceptibility-induced magneticfields, (c) fluid phases and their bulk properties, and (d)corresponding velocity fields superimposed to the diffusiveBrownian motion in the case of ionic displacement [17] orhydraulic motion [18]. After motivating the need for amicroscopic description of the pore space to constructquantitative NMR numerical models, we develop the algo-rithm for the solution of Bloch’s equations along the con-tinuous diffusion pathways of random walkers. Thisapproach remains valid when the magnetization couplingbetween spins in a fluid is governed by constant longitudi-nal and transversal bulk relaxations, and when Bloch’sequations remain valid to describe spin magnetization,i.e., both in the low-field NMR measurement of spin 1/2,and in the high-field 1H measurements of water. Ourapproach is benchmarked against analytical solutions forcomplex diffusion/relaxation problems. We then generalizethe application of the algorithm for simulating the NMR

response of fluids in porous media, and emphasize its keyrole in parametric multi-dimensional NMR modeling usingpore-scale arguments.

2. Macroscopic theory

Macroscopic descriptions of NMR measurements usuallyinvoke three independent processes to describe the relaxa-tion time of the measured magnetization: (1) bulk fluidrelaxation, with characteristic longitudinal and transversaltimes T1B and T2B, mainly due to dipole–dipole interac-tions between spins within the fluid; (2) surface relaxation,characterized by effective relaxation times T1S and T2S,which occurs because of the proximity with paramagneticcenters at interfaces between pore fluids and grain solids;and (3) relaxation due to the presence of background mag-netic field heterogeneities, T2D. The longitudinal and trans-versal relaxation times measured for a fluid-saturatedsample are equal to the harmonic averages [19]

1

T 1

¼ 1

T 1B

þ 1

T 1S

; ð1Þ

and

1

T 2

¼ 1

T 2B

þ 1

T 2S

þ 1

T 2D

: ð2Þ

Bulk and surface relaxation times are intrinsic propertiesgiven by fluid, porous medium, and thermodynamic state.Bulk relaxations are usually expressed in the form [20,21]

T 1B;2B ¼ aTg¼ f P ; Tð ÞDB; ð3Þ

where a is a fluid-specific constant, f a thermodynamicfunction, and P, T, g, and DB the pressure, temperature,viscosity, and bulk diffusivity of the fluid, respectively.When diffusion–relaxation is controlled by diffusion, i.e.,in the fast-diffusion limit qR/DB < 1 (R being the pore ra-dius), the surface decay times T1S and T2S are proportionalto S/V [1]

1

T 1S;2S

¼ q1;2

SV: ð4Þ

The variable T2D, on the other hand, is an experimentallycontrollable quantity affected by fluid diffusivity, internalfields induced by porous medium magnetic heterogeneities,and instrument configuration. For instance, using a CPMGsequence with constant echo time TE and an applied mag-netic gradient G, we have

1

T 2D

¼ c G TEð Þ2D12

; ð5Þ

where c is the hydrogen gyromagnetic ratio and D remainsthe effective diffusivity of the fluid under consideration.

The intrinsic bulk and surface relaxation times affect thetime-space evolution of the macroscopic magnetizationM = (Mx,My,Mz)

T of a given volume of fluid. Following

Page 3: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96 85

Bloch–Torrey’s equation in the rotating frameR ¼ ðx0; y0; zÞ, one has [22]:

dM

dt¼ DBr2Mþ cB�M�Mx

T 2

x0 �My

T 2

y0 �Mz

T 1

zþM0

T 1

:

ð6Þ

In this equation, M0 ¼ M0z is the equilibrium magnetiza-tion aligned with the background magnetic field B0, andB = (Bx,By,Bz)

T is the total magnetic field that affects thespins in R. The magnetic field B varies as a function of timeand space because of (a) inhomogeneous tool backgroundfield maps such as described in Fig. 1, (b) susceptibilitycontrasts between fluid and porous matrix which distortthe background field map [16,23], and (c) additional back-ground field gradients used to enhance the NMR decay dueto diffusivity contrasts between fluids in low-field applica-tions, or to encode the spin position in high-field MRIapplications. When a RF pulse sequence is used to enforcespin refocusing and echo forming, B1 magnetic pulsessuperimpose to B0 for finite durations th. This durationquantifies the amount of magnetization tilting (h = ciB1ith)

Fig. 1. Magnetic field map representing an approximation of thebackground magnetic field imposed by the measurement system andadapted from Ref. [12], assuming B0 of the elliptical form B0 =(�(x/c � 0.5)2 + (y/c � 0.5)2 · 30 z [Gauss], where c = 1 cm.

they create through the curl term in Eq. (6)—usually p/2 orp for the entire pulse duration. For instance, regularCPMG pulse sequences take the form of a p/2-pulse ofduration t90 along the x0 axis, followed by a series of ppulses of duration t180 along the y0 axis separated by the in-ter-echo time TE = 2s. Spin echoes are stimulated betweentwo successive p pulses. Together with the proximity ofrelaxing surfaces which affect the relaxation of the spinpackets captured by Eq. (6), this variation of B with bothtime and space supports a microscopic description whereinsuch constraints can be explicitly enforced in the numericalformulation.

3. Microscopic random-walk resolution

Random walkers reproducing the diffusive motion ofspins have proved to be a reliable and flexible approachto simulate the NMR responses in a variety of porousmedium geometries, including fractal or 3D disorderedgranular porous media [24,25]. In addition, diffusion ran-dom walks readily lend themselves to distributed com-puter environments, thereby considerably reducingcomputation time. Such algorithms simulate the projec-tion of spin packets in the transversal plane of the rotat-ing frame, i.e., in the plane of measurement of an NMRtool antenna, and describe either surface relaxation effects[4,26] or diffusion effects [23,27]. In this section, we intro-duce a unified formulation for the simulation of NMRmeasurements acquired for a wide range of engineeringand biological applications.

3.1. Diffusion of equivalent magnetic dipoles

We use a classical continuous random walk, whereby anequivalent magnetic dipole is displaced within the porespace attributed to a given fluid of self-diffusion DB andbulk relaxation times T1B and T2B. Following an initiallocation randomly determined within that fluid volume,at each time step of infinitesimal-duration dt a vector ofnormally distributed random numbers n = (nx,ny,nz) isgenerated with unit variance and zero mean. The walkeris then spatially displaced in the Cartesian geometrical ref-erence by a vector

drD ¼ ð6DB dtÞ1=2n=knk: ð7Þ

In the presence of velocity drift v applied to the molecule, anadditional drift displacement drV = vdt is added to drD. In sodoing, dt is dynamically adapted along the walk so that (a) itis smaller than any magnetic field pulse duration, (b) theamplitude of the total displacement, idr = drD + drVi, is sev-eral times smaller than surrounding geometric length scales(e.g., pore throats restricting pore-to-pore connections,wetting film thickness, etc.), (c) the value of the total mag-netic field resulting from B0 and B1 in the rotating framecan be assumed constant during each step, and (d) all walkers

Page 4: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

0200

400600

8001000 0

1

23

0

0.5

1

Wait time [s]

Acquisition time [ms]

Mag

netiz

atio

n

Mzpm

Mzbulk

Mypm for increasingTEs

Mybulk for increasingTEs

Fig. 2. Amplitude of the mean Mz and My magnetizations simulated for1-s CPMG pulse acquisitions taking place after different wait times for thesame volume of bulk water (dashed curves, notedbulk) and water relaxingin the porous medium (continuous curves, notedpm) formed by the voidfraction of the disordered grain packing shown in Fig. 10. Surfacerelaxivity (q1 = q2 = 20 lm/s) and pore size (30 lm) were considereduniform. Each magnetic dipole is initially depolarized (Mz = 0 and My

randomly distributed with zero mean), then Mz increases freely for a waittime TW (no RF pulse). Dipoles collectively generate an exponentialbuild-up of the form 1 � exp(�t/T1). The CPMG RF pulse sequence isthen turned on, Mz is tilted by the first (p/2)x-pulse into My which becomesnon-zero and subsequently decays. Next, dipoles collectively generate amacroscopic magnetization in the exponential form exp (�t/T2) followingthe subsequent (p)y refocusing pulses. The results plotted above weresimulated for TE = 0.3 ms at TW = 0.03, 0.1, 0.3, and 1 s, and forTE = 0.3, 1, 4, and 16 ms at TW = 3 s.

86 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

are synchronized at the onset and offset of magnetic pulses aswell as at the time of echo acquisition.

3.2. Magnetization formulation A

We use diffusing random walkers to simulate NMR sig-nals under the assumption that each walker carries a mag-netic-dipole equivalent to that of a hydrogen spin packet.Eq. (6) is adapted along each one-dimensional segmentby neglecting its convection term in $2 and by definingthe evolution matrix A given by

A ¼� 1

T 2cBz �cBy

�cBz � 1T 2

cBx

cBy �cBx � 1T 1

2664

3775; ð8Þ

and the reduced magnetization

M0 ¼Mþ A�1 �M0=T 1: ð9ÞEq. (6) is integrated as

M0ðtÞ ¼ eAt �M0ð0Þ; ð10Þwhere M 0(0) is the reduced magnetization at the onset ofeach integration interval. New appropriate values for Bx,By, and Bz are used on each random-walk segment. Therelaxation times T1 and T2 are also locally adjusted to ac-count for the proximity of a relaxing boundary. Conse-quently, for each step i spanning the time interval[ti, ti + dti], the spin carried by the random walker is subjectto a new evolution matrix Ai and longitudinal relaxationtime T i

1. This calculation yields the magnetization formula-tion A

Mi ¼ eðAiÞdti � Mi�1 þ Ai� ��1 �M0

T i1

� �� ðAiÞ�1 �M0

T i1

: ð11Þ

The total NMR signal is the sum of all M vectors for alldiffusing magnetic dipoles, sampled at the same rate.Hydrogen indices weigh the magnetization contributionof each walker in the presence of multiple fluid phases withdifferent hydrogen index. The Appendix provides analyticalsolutions for the exponential of matrix A. Tracking theprojections of that total magnetization M onto z and y0

with time yields the magnetization polarizations and de-cays illustrated in Fig. 2. This figure shows the responsesof both bulk water and water within relaxing porous med-ium, simulated with CPMG pulse sequences and differentvalues of wait times (TW) and echo time (TE). Fig. 3 illus-trates the joint influence of sampling and step size of therandom walks on the accuracy of the simulation results.

3.3. Magnetization formulation B

Previously developed NMR simulation algorithmsfocused exclusively on reproducing surface relaxation ordiffusion within inhomogeneous fields. The latter effectwas usually treated in a manner similar to the onedescribed above, except that Bloch’s equation was solved

in the ðx0; y0Þ projection of R by substituting M in Eq. (6)with a complex magnetization M* = Mx + iMy [28]. Inturn, the magnetization was calculated numerically usingfinite-difference methods [29] or without including a con-vection term along random-walk trajectories [23,27]. Thatapproach neglected surface as well as longitudinal relaxa-tion effects. Likewise, RF pulses were not accurately repro-duced, and (p)y pulses could only be treated as a signchange in the phase of M* [23], which made the methodexclusively amenable to CPMG sequences. The NMR sig-nal was then considered equal to the sum of the cosine ofthe spin phase shifts, where the phase shifts were equal tothe same rotation angle as in the evolution matrix for 3Dcalculations (see the Appendix), i.e., / = cBdt. For com-parison purposes, we also implemented this simulationmethod in ðx0; y0Þ space as formulation B.

3.4. Treatment of local surface relaxation

While the walker remains away from a fluid boundary,the relaxation times T1 and T2 that affect the equivalentmagnetic dipole of the random walker are equal to the bulkfluid longitudinal and transversal relaxation times, T1B andT2B, respectively. However, once the walker is locatedwithin one step of a fluid boundary of surface relaxivitiesq1 (longitudinal) and q2 (transversal), the magnetizationdecay is locally enhanced to include this surface relaxationeffect at the microscopic level. If the displacement achieved

Page 5: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

0 200 400 600 800 10000

0.2

0.4

0.6

0.8

1

Mea

n m

agne

tizat

ion

0 200 400 600 800 10000

0.2

0.4

0.6

0.8

1

Time [ms]

Mea

n m

agne

tizat

ion

AnalyticalN=200, δr = 1μmN=200, δr = 3μmN=200, δr = 30μm

AnalyticalN=2000, δr = 1μmN=2000, δr = 3μmN=2000, δr = 30μm

Fig. 3. Influence of walker number (N) and nominal step size (dr) on theaccuracy of the numerical simulations. This example simulates theresponse of a water volume of 2 · 2 · 2 mm, with G = 13.2 G/cm andTE = 2 ms. The resulting apparent T2 value is equal to 751 ms. Largevalues of N are necessary to obtain smooth decays. Even with large N

values, too large values of dr create a bias. In this example, no step sizelarger than 1 lm should be used for maximal accuracy.

E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96 87

during the interval [t, t + dt] comes within a distance dr of afluid boundary, the boundary equation

Dbulk

oM

onþ

q2

q2

q1

26664

37775 �M ¼ 0 ð12Þ

governs the behavior of the relaxation, where the unit vec-tor n describes the normal direction to the surface bound-ary. For a locally flat boundary, Eq. (12) is solved on asquare lattice [30] or for continuous random walks [26] todefine a probability of instantaneously ‘annihilating’ thespin when it reaches the boundary. Bergman et al. [26]solved the boundary condition by integrating diffusionpropagators and related that ‘annihilating’ probability p

to surface relaxivity q as

p ¼ 2

3

qdrDB

� 0:96: ð13Þ

In our model, we consider that each time the walkerreaches a relaxing boundary the magnetization decreasesby a factor (1 � p), or exp[�dt(p/dt)] since p� 1. There-fore, the total relaxation rate becomes 1/T1,2 = 1/T1B,2B +p/dt for that step. Equivalently, the relaxation times usedon each segment of the random walk become

1

T 1;2

¼ 1

T 1B;2B

þ e3:84q1;2

dr; ð14Þ

where e = 1 when the walker is located within a step of therelaxing boundary, and 0 otherwise.

3.5. Transfer probability of permeable membranes

The case of permeable barriers between pore volumescan be approached in a similar fashion to the case of relax-ing surfaces. A boundary exchange rate kj is defined as theproportionality factor between diffusion flux and concen-tration difference across a boundary j, in units of time�1.If bulk and surface relaxations can be neglected, thenthe macroscopic magnetization balance of a compart-ment bounded by permeable membranes is given in[7,8]

oM

ot¼ DBr2Mþ cB�M�

Xoutwards

kjMþX

inwards

kjMðjÞ;

ð15Þwhere M(j) is the magnetization within a neighboring com-partment j. The sum

Pover outward boundaries acts as a

surface sink term whereas the one over inward boundariesacts as a surface source term. Since the surface sink term inEq. (12) acts as the relaxation decay terms in Eq. (1), thefollowing equivalence holds:

1

T 2

�X

outwards

kj: ð16Þ

Furthermore, in the fast-diffusion limit we have 1/T2S =q(S/V) [1]; similarly, kj = Pj(S/V) [7], where Pj is the mem-brane permeability in units of length per unit time. It isthen possible to solve for surface permeability usingBergman et al.’s treatment of surface relaxation: whenthe walker comes within one step of distance dr from a per-meable boundary k, there is a probability of transfer equalto

gk ¼2

3

P kdrDB

� 0:96 ð17Þ

that the walker crosses the boundary; if that probability isnot honored, then the walker bounces back into its originalcompartment while the clock time is incremented by dt. Itis particularly remarkable that both fluid/rock surfacerelaxation [31] and biological membrane permeability [7]exhibit the same range of values between 0.5 and 30 lm/s.This similarity suggests that the parallel treatment ofrelaxing or permeable boundaries remains valid on thesame range of fluid types and pore sizes.

4. Simulation results in bulk fluid

4.1. Echo sampling

The accuracy of our algorithm is first tested for thecase of a homogeneous background magnetic field

Page 6: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

88 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

gradient. Fig. 4 shows the shapes of the 1st, 2nd, 5th,and 15th CPMG echoes simulated with the random-walk algorithm for different echo times. The resultsare in excellent agreement with both the measurementsand the coherent-pathway calculations of Hurlimann inRef. [13]. We can therefore conclude that random-walksimulations are a viable and easily implementedalternative to the more complex coherent-pathwaycalculations.

Measurements

Sig

nal

–100 –50 0 50 100

1.2

0.8

0.4

0.0

Sig

nal

–100 –50 0 50 100

1.2

0.8

0.4

0.0

Sig

nal

–100 –50 0 50 100

1.2

0.8

0.4

0.0

Time [μs]

Sig

nal

–100 –50 0 50 100

1.2

0.8

0.4

0.0

Tim–100 –50

1.2

0.8

0.4

0.0

–100 –50

1.2

0.8

0.4

0.0

–100 –50

1.2

0.8

0.4

0.0

Coherentcalcu

–100 –50

1.2

0.8

0.4

0.01st

2nd

5th

15th15th echo

5th echo

2nd echo

1st echo

Fig. 4. Comparison between measurements and coherent-pathway calculationand 15th echoes obtained for a CPMG sequence with a water sample. The rand2 · 2 · 2 cm. All results use the same parameters: homogeneous gradient of strecho, TE = 0.4, 8.35, 10.35, 12.35, 14.35, and 16.35 ms; for the 2nd echo, TE =3.75, 4.95, 6.35, and 8.35 ms; for the 15th echo, TE = 0.4, 1.354, 2.75, 3.75, 4.pulses that immediately precede and follow the echoes.

Hurlimann and Griffin [12] quantified the relative ampli-tude variation of the stimulated echo (produced by theenergy stored during the initial ðp=2Þx0-pulse) and of thefirst spin echo (refocused by the first CPMG ðpÞy0 pulses)with coherent-pathway calculations for given field mapsand pulse durations. In Fig. 5 we show that similar resultsare obtained with the random-walk technique. The differ-ences between the results of Ref. [12] and our simulationscan be attributed to differences in field map details and

e [μs]0 50 100

0 50 100

0 50 100

pathway lations

0 50 100 –100 –50 0 50 100

0.0

0.4

0.8

1.2

Random walk calculations

–100 –50 0 50 100

0.0

0.4

0.8

1.2

–100 –50 0 50 100

0.0

0.4

0.8

1.2

–100 –50 0 50 100

0.0

0.4

0.8

1.2

Time [μs]

echo

echo

echo

echo 15th echo

5th echo

2nd echo

1st echo

s from Hurlimann [13], and random-walk simulations of the 1st, 2nd, 5th,om-walk simulations considered 20,000 walkers distributed in a volume ofength G = 13.2 G/cm; pulse durations t180 = 50 ls, t90 = 25 ls; for the 1st0.4, 4.35, 6.35, 8.35, 10.35, and 12.35 ms; for the 5th echo, TE = 0.4, 2.35,

95, and 6.35 ms. The time scale is centered with the midpoint between the

Page 7: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

2 4 6 8 10 12 14 160

0.2

0.4

0.6

0.8

1

Echo time [ms]

Ech

o am

plitu

de

Fig. 5. Amplitudes of the first (dash) and second (plain) CPMG echoessimulated as a function of echo time for 1 cm3 of bulk water volume.Curves with square and circle markers identify the mean My signalsimulated with 10,000 random walkers in the field map of Fig. 1, fordifferent B1 pulse widths (square markers: t90 = 10 ls, t180 = 15 ls; circlemarkers: t90 = 20 ls, t180 = 30 ls). Calculations reported in Ref. [12] for asimilar field map are plotted with open triangle markers.

E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96 89

homogeneity of the background magnetic field. In all cases,the amplitude of the first echo is remarkably lower thanthat of the second echo for TE <5 ms, and becomes largerwhen TE increases.

Dispersion of magnetic field strength—and therefore ofLarmor frequencies—throughout the volume enhances theprecession decay and the creation of echoes due to the RFp pulses of a CPMG sequence. When the duration of theseRF pulses increases, then the measurable echo amplitude

0 .3 .6 .9 1.2 1.5 1.8

0

0.2

0.4

0.6

0.8

1

Mea

n m

agne

tizat

ion

t90

= 10 μs , t180

= 15 μs

0 .3 .6 .

0

0.2

0.4

0.6

0.8

1

0 .3 .6 .9 1.2 1.5 1.8

0

0.2

0.4

0.6

0.8

1

Mea

n m

agne

tizat

ion

0 .3 .6 .

0

0.2

0.4

0.6

0.8

1

Tim

homogeneous gradient

approximate logging tool map approximate

homogen

t90

= 20 μs

Time [ms]

Fig. 6. Impact of background gradient on NMR echo amplitudes. The mean N1 cm3, 300-ls echo time CPMG sequence, and different values of t90 and t180. Pnoise); dashed curves, Mz. Top row: simulations performed with a homogeneowith the field map described in Fig. 1.

decreases. Background field maps can therefore be opti-mized to maximize the signal obtained during measurementacquisition. These remarks are illustrated in Fig. 6, wherethe formation of echoes in 1 cm3 of water is simulated withour random-walk method (formulation A) through spinrefocusing between two p pulses, for different pulse widthsand two permanent magnetic field maps. As the duration ofthe RF pulses increases, the quality of the echoes degradesthrough lower amplitudes and larger spread. In these tests,the field map shown in Fig. 1 generates higher echo ampli-tudes, hence higher signal, than that of a homogeneousmagnetic gradient.

The magnetization decays considered in the followingsections of this paper are taken as the envelopes of maximalvalues reached at each My echo.

4.2. Decay in an inhomogeneous background magnetic field

We now compare the decays generated by the presenceof background magnetic gradients between (1) random-walk formulation A, (2) random-walk formulation B, and(3) the normalized apparent decay M ¼ exp½�tð1=T 2Bþ1=T 2DÞ�. Because the My signal obtained using formulationA is sensitive to large sampling volumes and imperfectionsin the magnetic field distribution, all its simulated decayswere normalized at the time origin. Fig. 7 compares simu-lation results for two values of homogeneous gradient,Gz = 12 and 29 G/cm, three echo times, TE = 0.3, 1, and3 ms, and two sampling volumes, 1 mm3 and 1 cm3. For-mulation B yields the same decays regardless of the sample

9 1.2 1.5 1.8 0 .3 .6 .9 1.2 1.5 1.8

0

0.2

0.4

0.6

0.8

1

t90

= 30 μs , t180

= 45 μs

9 1.2 1.5 1.8e [ms]

0 .3 .6 .9 1.2 1.5 1.8

0

0.2

0.4

0.6

0.8

1

Time [ms]

logging tool map approximate logging tool map

eous gradient homogeneous gradient

, t180

= 30 μs

MR magnetization is calculated from 20,000 walkers for a water volume oflain curves identify My projection (signal); dotted curves, Mx (out-of-phaseus permanent gradient Gz = 20 G/cm. Bottom row: simulations performed

Page 8: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

0 .1 .2 .3 .4 .5 .6 .7 .8 .9 10

0.2

0.4

0.6

0.8

1.0

Time [s]

Nor

mal

ized

am

plitu

de

Gz = 12 G/cm

0 .1 .2 .3 .4 .5 .6 .7 .8 .9 10

0.2

0.4

0.6

0.8

1.0

Time [s]

Gz = 29 G/cm

TE = 0.3 ms

TE = 1 ms

TE = 3 ms

TE = 0.3 ms

TE = 1 ms

TE = 3 ms Nor

mal

ized

am

plitu

deFig. 7. NMR decay for CPMG pulse acquisition of bulk water in a homogeneous background magnetic gradient Gz for different values of TE and formagnetization formulations A and B. Open square markers describe the analytical solution; simulation results of Formulation A are described bycontinuous curves for a 1 mm3 simulated water sample, and by dots for a 1 cm3 water sample. For each value of TE and each sampling volume,Formulation B yields results that exactly match the analytical solution.

0 0.5 1 1.5 2

0

.2

.4

.6

.8

1

Am

plitu

de

0 .2 .4 .6 .8 1

0

.2

.4

.6

.8

1 TW = 1 msTW = 10 msTW = 100 msTW = 1 sTW = 10 s

7–cp Oilunimodal

T2B

= 200 ms

90 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

volume and exactly matches the analytical decays. Formu-lation A also perfectly fits the analytical results for thesmaller sampling volume, but exhibits substantial differ-ences for the larger volume. This effect is due to the distri-bution of the energy acquired during the different RFpulses, as mentioned in Section 4.1, and cannot beaccounted for by formulation B or by standard analyticalexpressions that disregard the finite duration of the RFpulses.

0 0.5 1 1.5 2

0

.2

.4

.6

.8

1

Time [s]

Am

plitu

de

0 .2 .4 .6 .8 1

0

.2

.4

.6

.8

1

Time [s]

10–1

100

101

102

T2B

Relaxation time [ms]

Am

plitu

de300–cp Oilmultimodal

T2Bpeak = 40 ms

Fig. 8. NMR magnetization decays of bulk fluids simulated for differentvalues of wait time (left panels), and equivalent decays normalized inamplitude at time t = 0 (right-hand panels). The same legend applies to allpanels. For simple fluids with unimodal distribution of bulk relaxationtimes (top-right panel), all curves scale with TW and overlap with thecanonical form (1 � exp(�TW/T1)) · exp(�t/T2). For multi-componenthydrocarbons (bottom right-hand panel), such normalized curves do notoverlap. The insert describes the distribution of bulk relaxation times forthe 300-cp heavy oil measured at laboratory conditions, and is used asinput to the simulation algorithm.

4.3. T1 scaling

It is customary to consider that the NMR decayobtained for a fluid after partial repolarization of durationTW is equal to the decay at full polarization (i.e., startingfrom M = M0, or TW fi1) weighted by the value(1 � exp(�TW/T1)). Fig. 8 shows that this treatment isappropriate for a simple-component fluid such as 7-cp oilwith unimodal bulk relaxation time at 200-ms in laboratoryconditions. For a 300-cp multi-component heavy oil with40-ms relaxation peak value (distribution shown in theinsert of Fig. 8), the decays obtained after partial polariza-tion are normalized so that the amplitude of each decay is 1when t = 0. For a known distribution of bulk relaxationsof the multi-component fluid, each walker accounted fora molecule of one miscible component of that fluid, andwas stochastically assigned values of bulk relaxation timeand diffusivity so that the T2 distribution was honored overthe entire population of random walkers. The normalizeddecays show that no scaling exists between the differentdecays due to the wide distribution of relaxation times.In the presence of strong internal fields, the apparent valuesof T2 would also span a wide relaxation range, but without

the corresponding spread in the T1 spectrum. Again, in thiscase no scaling will exist between the magnetization decaysmeasured after incomplete polarizations with differentwait times. Because only Formulation A incorporates T1

build-up, it needs to be used for the cases of strong

Page 9: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

0 10 20 30 40 50 60 10

0

101

102

Mag

netiz

atio

n

0 20 40 60 80 10010

0

101

102

Mag

netiz

atio

n

0 100 200 300 400 500 10

0

101

102

Mag

netiz

atio

n

0 200 400 600 800 1000

102

Time [μs]

Mag

netiz

atio

n

R=0.4µm

R=4µm

R=40µm

R=400µm

(ρR)/DB=[4/20/200μm/s*400μm / 2.5μm2/ms]=0.64/3.2/32

(ρR)/DB=[4/20/200μm/s*40μm / 2.5μm2/ms]=0.064/0.32/3.2

(ρR)/DB=[4/20/200μm/s*4μm / 2.5μm2/ms]=0.0064/0.032/0.32

(ρR)/DB=[4/20/200μm/s*.4μm / 2.5μm2/ms]=0.00064/0.0032/0.032

Fig. 9. Decay curves simulated for water (DB = 2.5 lm2/ms, T2B = 3 s) ina single spherical pore with no background magnetic field gradient. Thepore radius R varies logarithmically from 0.4 (top panel) to 400 lm(bottom panel), while for each radius the surface relaxivity q at the porewall is equal to 4 lm/s (triangles), 20 lm/s (circles) or 200 lm/s (squares).These markers define the first-order analytical decay for a sphere(1/T2 = 1/T2B + 3q/R) which diverges from the full expression of Brownsteinand Tarr [1] (in dashes) when qR/DB reaches 1, Eq. (18). Results simulatedwith formulations A and B overlap (plain lines) and almost perfectlymatch the full analytical expression in all cases. The ratio qR/DB iscomputed within the panels for each combination of R and q.

E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96 91

heterogeneities in bulk relaxation and internal fields toensure accurate simulation results for partial polarization.

5. Simulation results in saturated porous media

We now study how the incorporation of a porous med-ium exhibiting surface relaxation affects the quality of thesimulation results. Different relaxation types are incorpo-rated simultaneously to generate parametric multi-dimen-sional NMR maps with the purpose of characterizingpore size, wettability, fluid type and volumetric fractionin the context of saturated rock-core analysis and in-situoil field evaluation.

5.1. Surface relaxation in different diffusion/relaxation

regimes

We test our algorithm for the classical context of an iso-lated spherical pore of radius R, where S/V = 3/R. Fig. 9shows the results of simulations performed with differentspherical pore radii ranging from 0.4 to 400 lm, and fordifferent values of surface relaxivity. In all the cases, thesimulation decays obtained with both formulations Aand B perfectly overlap in the absence of q1/q2 contrast,as encountered in natural systems; for example, it isobserved that this contrast usually does not exceed 1.5 inwater-saturated rocks [31]. If this contrast were to reachseveral units, however, differences in apparent decay timeswould appear between formulations A and B. Likewise, alocal increase of the gradient due to diamagnetic or para-magnetic materials in the solid phase of the porous mediumis more likely to affect simulation results using formulationA because of the finite duration of the B1 pulses.

The simulation results shown in Fig. 9 match the first-order surface relaxation time described by Eq. (4) as longas qR/DB < 1. Identical results were obtained in the 3Dstar-shaped pores formed by the void space fraction exist-ing within a dense packing of overlapping spherical grainswith different compaction coefficients.

Outside the fast-diffusion limit, that is when qR/DB

becomes larger than 1, the surface relaxation describedby Eq. (4) becomes inaccurate. It is then necessary to con-sider Brownstein and Tarr’s [1] full analytical solution,whereby the total magnetization decay due to surface relax-ation is equal to

MðtÞ ¼ Mð0ÞX1n¼0

Ine�t=T n : ð18Þ

In this expression, In ¼ 12 sin fn�fn cos fnð Þ2

f3n 2fn�sin 2fnð Þ½ � and T n ¼ R2=Df2

n are

the amplitude and relaxation time defined by fn, the nth po-sitive root of the equation

1� fn cot fn ¼ qR=D: ð19ÞIn Fig. 9, simulation results show the deviation from Eq.(4) to Eq. (18) is correctly captured by the random-walkalgorithm for any value of qR/DB.

5.2. Two-dimensional NMR maps of gas/water and oil/water

mixtures in saturated rocks

In oil exploration and rock/fluid characterization, fluidsmeasured by in-situ NMR logging tools include brine, sin-gle- and multi-component oil, gas, and water-based or oil-based drilling-mud filtrate which leaked into a freshlydrilled porous rock formation. Multi-dimensional NMRtechniques are currently based on the simultaneous

Page 10: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

Fig. 10. Example of disordered packing of identical grains used as rockmodel to generate the 2D NMR maps shown in Figs. 12 and 13. Thespherical grains are overgrown until the void fraction reaches 20% of thebulk volume.

92 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

inversion of suites of CPMG-like sequences with multiplevalues of TE and/or TW which allow different degrees ofspin repolarization between successive CPMG RFsequences (Fig. 2). Fluid phases are distributed in the porespace of a disordered grain pack (Fig. 10) according to geo-metrical models in agreement with thermodynamic capil-lary equilibrium [32] or with more complex principles ofdrainage, imbibition, and electrical properties [33]. In thesemodels, as exemplified in Fig. 11, fluid phases are distrib-uted at the pore level as overlapping blobs which are trun-cated by grain surfaces. Two examples of rock/fluidsystems were simulated to illustrate the methodology indisordered packs of overlapping spheres exhibiting 20%void space, uniform distribution of pore sizes, and enforc-ing a homogeneous gradient with Gz = 16 G/cm.

The first example was simulated for a mixture of 30%water and 70% gas in arbitrary reservoir conditions, whosebulk properties are shown in Table 1. Wait times varyinglogarithmically from 1 ms to 10 s were applied during thesimulations of CPMG decay with 0.2 ms echo time, andthe corresponding magnetization decays were processedwith a T1/T2 inversion algorithm [11]. The left-hand panelof Fig. 12 shows the resulting intensity T1/T2 map, wherethe water peak located at T1 = T2 = 100 ms is affected bysurface decay. The small echo time has negligible influence

Table 1Bulk properties of the fluids used in the numerical simulation of NMRmeasurements of saturated rocks

Fluid Hydrogenindex

Bulk relaxation timesT1B = T2B

Bulk diffusivityDB (cm2/s)

Water 1.0 3 s 2 Æ 10�5

Gas 0.3 4.5 s 10�3

7-cp oil 1.0 0.2 ms 10�6

on the T2 response of the low-diffusivity water; however, itsignificantly reduces the apparent T2 value of the large-dif-fusivity gas with respect to its intrinsic T1 value. In thiscase, inversion of the T1 dimension efficiently discriminatesthe signals from the two fluids. We also synthesized magne-tization decays for multiple-TE diffusion sequences for D/T2

inversion and mapping. This was done by considering fullpolarization (TW fi1) and values of TE equal to0.2, 1, 3, 9, and 16 ms for a homogeneous gradientGz = 16 G/cm. Focusing on the same water/gas mixturesaturating our porous rock model, we obtain the D/T2

map shown in the right-hand panel of Fig. 12. In this plot,as in subsequent D/T2 maps, the diagonal line characterizesa D/T2 correlation existing for live hydrocarbons at specificconditions of pressure and temperature (Eq. (3)). In thisplot, water and gas are distinguishable on both the D

and T2 dimensions.Our second example consists of a mixture of 60% water

and 40% 7-cp oil at ambient conditions (see Table 1) withinthe same rock model, and illustrates the simulation of wet-tability effects at the rock surface. The same diffusionsequences were implemented, and the simulation resultswere inverted into the D/T2 maps shown in Fig. 13. Simula-tion results for the water-wet case (left-hand panel) exhibit awater peak at the intersection of the bulk water diffusivity(2.105 m2/ms) and the surface relaxation time previouslyobserved in Fig. 13 (100 ms). The oil peak, however, liesalong the D/T2 hydrocarbon correlation line. The (D,T2)coordinates of that peak along that line enable the assess-ment of oil viscosity and possibly water-wettability. Uponwettability alteration, it is usually assumed that the waterfilm existing between oil and rock surface ruptures underdifferent chemical and thermodynamic conditions [34,35],while the least attainable zones of the pore space remainwater-wet (see Fig. 11). The right-hand panel of Fig. 13shows simulation results obtained when such microscopicgeometries are included in the model. When oil is consideredas the wetting fluid phase, it is affected by surface relaxationand the spread of its D/T2 peak increases along both the D

and T2 dimensions. Concomitantly, the relaxation time ofthe water peak increases because some relaxing water/rocksurfaces have disappeared, and some of the water nowrelaxes in bulk mode. Because of the low T1B/T2B contrastsfor water and oil, the low gradient value, and because of thechoice of CPMG acquisition sequence, Formulation B wascomputationally efficient in this particular case. We empha-size, however, that Formulation A would be required formore general RF sequences or in the presence of strongermagnetic field gradients.

The random-walk algorithm developed in this paper canbe used to quantify the impact on NMR measurements ofthe combined effects of fluid content, fluid viscosity, satura-tion history of the porous medium, and rock wettability.For instance, Refs. [32,33] study the reliability of D/T2

2D NMR interpretation to accurately assess hydrocarboncontent and rock properties under the assumption of spe-cific fluid displacement mechanisms within the pore space.

Page 11: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

Fig. 12. T1/T2 (left) and D/T2 (right) NMR maps simulated for a two-phase immiscible mixture of 70% gas and 30% water in the grain packing of Fig. 10.The diagonal line in the left-hand panel represents the T1 = T2 relationship. The diagonal line in the right-hand panel represents the D/T2 correlationcharacterizing hydrocarbons in the conditions assumed for the simulations (after Ref. [32]).

Fig. 11. Examples of fluid distributions implemented at the pore scale. Each tetrahedron is centered on a pore from the void space of the grain packing(Fig. 10), and is limited by sets of four-closest grains (in gray). Blue represents the water-filled pore space, red the oil-filled pore space. (a) Water fills thepore. (b) An oil blob centered with the pore under the double assumption of oil-wettability (OW) of the grain surface within the radius of the oil blob, andwater-wettability (WW) of the grain surface in the least-accessible pore regions. (c) A thin film of wetting oil and bridging oil lenses are left in the oil-wetregion of the pore after invasion by water. The random-walk step is adjusted within each fluid zone.

Fig. 13. D/T2 NMR maps simulated for a two-phase immiscible mixture of 60% water and 40% 7-cp oil in the grain packing of Fig. 10, for water-wet (left-hand panel) and oil-wet (right-hand panel) configurations. The diagonal line in the two panels represents the D/T2 correlation characterizing hydrocarbonsin the conditions assumed for the simulations (after Ref. [32]).

E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96 93

Page 12: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

94 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

6. Conclusions

The simulation algorithm described in this paper consol-idates existing NMR numerical simulation methods used indiverse areas of biology and engineering into a single for-mulation that incorporates both diffusion within inhomo-geneous magnetic fields and surface effects, includingmagnetization relaxivity and membrane permeability.Other than the magnetic-dipole approximation, this formu-lation is constrained by none of the assumptions requiredfor analytical models concerning the statistical distributionof spin phase shifts, spin density, field homogeneity, orfast-diffusion limit. In addition, the simulation algorithmdescribed in this paper extends NMR simulation capabili-ties beyond those of single-phase periodic media, whichare so far the only type of porous media that can be accu-rately described with analytical techniques.

Previously described NMR simulations techniquesneglected longitudinal effects; while it is expected that thosetechniques would not accurately reproduce the NMRresponse of molecules with substantial T1/T2 contrasts,we showed that the solution of the complete set of Bloch’sequations is necessary to accurately reproduce echo ampli-tudes in the presence of strongly inhomogeneous magneticfields and arbitrary pulse sequences. The simulation algo-rithm described in this paper was successfully validatedagainst coherent-pathway results in homogeneous gradi-ents and against the full expansion of Brownstein andTarr’s surface relaxation in spherical pores; these resultssuggest that random walkers can reliably be used to accu-rately characterize diffusion in arbitrary porous media,where analytical expressions are not accurate.

We considered suites of longitudinal and transversalmagnetizations for multiphase fluid mixtures saturatingmodels of porous media for different values of wait timeand echo time. These synthetic magnetizations can besubsequently inverted into parametric multi-dimensionalintensity maps cross-plotting T1, T2, and D for improvedcharacterization of pore structure, wettability, and fluidtypes. Such a methodology provides new opportunitiesto study quantitative relationships between NMR mea-surements and properties of economic interest thatcannot be directly measured in-situ [31]. Moreover, theaccurate simulation of magnetization decay opens thepossibility of optimizing new NMR pulse acquisitionsequences for specific applications in porous media inthe presence of local variations of fluid distributionsand internal fields.

Appendix A. Analytical calculations of the time-evolution

matrix exponentials in Bloch’s equation

The total magnetic field applied to the spin B = B0 + B1

is either B ¼ B0 ¼ B0z between RF pulses, B ¼ B1xþ B0z

during x 0-pulses, or B ¼ B1yþ B0z during y 0-pulses. Thisapproach makes no approximation on the quality of mag-netic pulses because the actual dispersion of Larmor fre-

quencies is honored across the spin population and thespin diffusion is fully accounted for during the enforcementof B1.

Between B1 pulses, the only magnetic field present is thebackground field, and the effective relaxation times T1,2 forthat step are given by Eq. (14). It then follows that

Adt ¼� dt

T 2cBzdt 0

�cBzdt � dtT 2

0

0 0 � DtT 1

2664

3775; and therefore

eAdt ¼e�

dtT 2 cos cBzdtð Þ e�

dtT 2 sin cBzdtð Þ 0

�e� dt

T 2 sin cBzdtð Þ e� dt

T 2 cos cBzdtð Þ 0

0 0 e�dtT 1

2664

3775: ðA1Þ

During B1 (p/2)x pulses, the product Adt takes on the form

Adt ¼� dt

T 2cBzdt 0

�cBzdt � dtT 2

cBxdt

0 �cBxdt � dtT 1

2664

3775: ðA2Þ

The three eigenvalues ki of Adt are found by solving thecharacteristic polynomial given by

P xðkÞ ¼ detðkI� AdtÞ

¼ k3 þ k2 2dtT 2

þ dtT 1

� �

þ k2ðdtÞ2

T 2T 1

þ ðdtÞ2

T 22

þ ðcBxdtÞ2 þ ðcBzdtÞ2 !

þ ðdtÞ3

T 22T 1

þ ðdtÞ3ðcBxÞ2

T 2

þ ðdtÞ3ðcBzÞ2

T 1

!;

which we solve using Cardan’s method

k1 ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðs� qÞ3

p�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðsþ qÞ3

p� a

3;

k2 ¼ 12

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðsþ qÞ3

p�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðs� qÞ3

p� �� a

3þ i

ffiffi3p

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðsþ qÞ3

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðs� qÞ3

p� �;

k3 ¼ 12

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðsþ qÞ3

p�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðs� qÞ3

p� �� a

3� i

ffiffi3p

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðsþ qÞ3

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiðs� qÞ3

p� �;

8>>>><>>>>:

ðA3Þ

with p ¼ 1

3b� a2

3

� �; q ¼ 1

2c� ab

3þ 2a3

27

� �;

s ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiq2 þ p3

p; and

a ¼ 2dtT 2

þ dtT 1

� �; b ¼ 2 dtð Þ2

T 2T 1

þ dtð Þ2

T 22

þ cBxdtð Þ2 þ cBzdtð Þ2 !

;

c ¼ dtð Þ3

T 22T 1

þ dtð Þ3 cBxð Þ2

T 2

þ dtð Þ3 cBzð Þ2

T 1

!:

ðA4Þ

A matrix of eigenvectors corresponding to the above eigen-values can be calculated as

Page 13: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96 95

U ¼

cBzdtk1þ dt

T 2

cBzdtk2þ dt

T 2

cBzdtk3þ dt

T 2

1 1 1

� cBxdtk1þ dt

T 1

� cBxdtk2þ dt

T 1

� cBxdtk3þ dt

T 1

2666666664

3777777775;

whereupon,

eAdt ¼ U �ek1 0 0

0 ek2 0

0 0 ek3

264

375 �U�1:

During B1 (p)y pulses, Adt takes on the form

Adt ¼� dt

T 2cBzdt �cBydt

�cBzdt � dtT 2

0

cBydt 0 � dtT 1

2664

3775:

Upon replacing Bx with By in Eq. (A4), it is readily foundthat the eigenvalues of this last matrix are identical to thoseof Eq. (A3). The corresponding eigenvectors are includedin matrix U as

U ¼

1 1 1

� cBzdtk1þ dt

T 2

� cBzdtk2þ dt

T 2

� cBzdtk3þ dt

T 2

cBydtk1þ dt

T 1

cBydtk2þ dt

T 1

cBydtk3þ dt

T 1

266666664

377777775:

The computation of eAdt follows directly from the aboveexpressions.

References

[1] K.R. Brownstein, C.E. Tarr, Importance of classical diffusion inNMR studies of water in biological cells, Phys. Rev. A 19 (1979)2446–2453.

[2] L.L. Latour, P.P. Mitra, R.L. Kleinberg, C.H. Sotak, Time-depen-dent diffusion coefficient of fluids in porous media as a probe ofsurface-to-volume ratio, J. Magn. Res. A 101 (1993) 342–346.

[3] Y.-Q. Song, Determining pore sizes using an internal magnetic field,J. Magn. Res. 143 (2000) 397–401.

[4] T.S. Ramakrishnan, L.M. Schwartz, E.J. Fordham, W.E. Kenyon,D.J. Wilkinson, Forward models for nuclear magnetic resonance incarbonate rocks, Log Analyst 40 (1999) 260–270.

[5] E. von Meerwall, R.D. Ferguson, Interpreting pulsed-gradient spin-echo diffusion experiments with permeable membranes, J. Chem.Phys. 74 (1981) 6956–6959.

[6] J. Pfeuffer, W. Dreher, E. Sykova, D. Leibfritz, Water signalattenuation in diffusion-weighted 1H NMR experiments duringcerebral ischemia: influence of intracellular restrictions, extracellu-lar tortuosity, and exchange, Magn. Res. Imag. 16 (1998) 1023–1032.

[7] G.J. Stanisz, A. Szafer, G.A. Wright, R.M. Henkelman, An analyticalmodel of restricted diffusion in bovine optic nerve, Magn. Res. Med.37 (1997) 103–111.

[8] G.P. Zientara, J.H. Freed, Spin-echoes for diffusion in bounded,heterogeneous media: a numerical study, J. Chem. Phys. 72 (1980)1285–1292.

[9] M.D. Hurlimann, L. Venkataramanan, Quantitative measurement oftwo-dimensional distribution functions of diffusion and relaxation ingrossly inhomogeneous fields, J. Magn. Res. 157 (2002) 31–42.

[10] B. Sun, K.-J. Dunn, Probing the internal field gradients of porousmedia, Phys. Rev. E 65 (2002) 051309.

[11] B. Sun, K.-J. Dunn, A global inversion method for multi-dimensionalNMR logging, J. Magn. Res. 172 (2005) 152–160.

[12] M.D. Hurlimann, D.D. Griffin, Spin dynamics of Carr–Purcell–Meiboom–Gill-like sequences in grossly inhomogeneous B0 and B1

fields and application to NMR well logging, J. Magn. Res. 143 (2000)120–135.

[13] M.D. Hurlimann, Diffusion and relaxation effects in general strayfield NMR experiments, J. Magn. Res. 148 (2001) 367–378.

[14] Y.-Q. Song, Categories of coherence pathways for the CPMGsequence, J. Magn. Res. 157 (2002) 82–91.

[15] Y. Zur, An algorithm to calculate the NMR signal of a multi spin-echo sequence with relaxation and spin-diffusion, J. Magn. Res. 171(2004) 97–106.

[16] K.-J. Dunn, Enhanced transverse relaxation in porous media due tointernal field gradients, J. Magn. Res. 156 (2002) 171–180.

[17] C.S. Johnson Jr., Q. He, Electrophoretic nuclear magnetic resonance,in: J.S. Waugh (Ed.), Advances in Magnetic Resonance 13, AcademicPress, London, 1989, pp. 131–159.

[18] A. Caprihan, E. Fukushima, Flow measurements by NMR, Phys.Rep. 198 (1990) 195–235.

[19] E. Fukushima, S.B.W. Roeder, Experimental Pulse NMR: ANuts and Bolts Approach, Addison-Wesley, Reading, Massachu-setts, 1981.

[20] S.-W. Lo, G.J. Hirasaki,W.V. House, R. Kobayashi, Mixing rulesand correlations of NMR relaxation time with viscosity, diffusivity,and gas/oil ratio of methane/hydrocarbon mixtures, SPE paper77264, SPE J. 7 (2002) 24–34.

[21] M. Winkler, J.J. Freeman, M. Appel, The limits of fluid propertycorrelations used in NMR well-logging: an experimental study ofreservoir fluids at reservoir conditions, Petrophysics 46 (2005) 104–112.

[22] H. Torrey, Bloch equations with diffusion terms, Phys. Rev. 104(1956) 563–565.

[23] R.M. Weiskoff, C.S. Zuo, J.L. Boxerman, B.R. Rosen, Microscopicsusceptibility variation and transverse relaxation: theory and exper-iment, Magn. Res. Med. 31 (1994) 601–609.

[24] L.M. Schwartz, J.R. Banavar, Transport properties of disorderedcontinuum systems, Phys. Rev. B 39 (1989) 11965–11970.

[25] D.J. Wilkinson, D.L. Johnson, L.M. Schwartz, Nuclear magneticrelaxation in porous media: the role of the mean lifetime s (q,D),Phys. Rev. B 44 (1991) 4960–4973.

[26] D.J. Bergman, K.-J. Dunn, L.M. Schwartz, P.P. Mitra, Self-diffusionin a periodic porous medium: a comparison of different approaches,Phys. Rev. E 51 (1995) 3393–3400.

[27] H. Gudbjartsson, S. Patz, NMR diffusion simulation based onconditional random walk, IEEE Trans. Med. Imag. 14 (1995)636–642.

[28] C.P. Slichter, Principles of Magnetic Resonance, Springer, New York,1992.

[29] P. Sen, A. Andre, S. Axelrod, Spin echoes of nuclear magnetizationdiffusing in a constant magnetic field gradient and in a restrictedgeometry, J. Chem. Phys. 111 (1999) 6548–6555.

[30] K.S. Mendelson, Continuum and random-walk models of mag-netic relaxation in porous media, Phys. Rev. B 47 (1993)1081–1083.

[31] W.E. Kenyon, Petrophysical principles of applications of NMRlogging, Log Analyst 38 (1997) 21–43.

[32] E. Toumelin, C. Torres-Verdın, B. Sun, K.-J. Dunn, Limits of2D NMR interpretation techniques to quantify pore size,wettability, and fluid type: a numerical sensitivity study, SPE J.11 (2006) 354–363.

[33] E. Toumelin, C. Torres-Verdın, S. Devarajan, B. Sun, Anintegrated pore-scale approach for the simulation of grain

Page 14: Random-walk technique for simulating NMR measurements and ... · 90 along the x^0 axis, followed by a series of p pulses of duration t 180 along the y^0 axis separated by the in-ter-echo

96 E. Toumelin et al. / Journal of Magnetic Resonance 188 (2007) 83–96

morphology, wettability, and saturation-history effects on elec-trical resistivity and NMR measurements of saturated rocks,Paper presented at the 2006 International Symposium of theSociety of Core Analysts, Trondheim, Norway, September12–16.

[34] G. Hirasaki, Wettability: fundamentals and surface forces, SPEFormation Eval. 6 (1991) 217–226.

[35] A.R. Kovscek, H. Wong, C.J. Radke, A pore-level scenario for thedevelopment of mixed wettability in oil reservoirs, AIChE J. 39 (1993)1072–1085.