Nuclear magnetic resonance in porous media

Last updated

Nuclear magnetic resonance (NMR) in porous materials covers the application of using NMR as a tool to study the structure of porous media and various processes occurring in them. [1] This technique allows the determination of characteristics such as the porosity and pore size distribution, the permeability, the water saturation, the wettability, etc.

Contents

Theory of relaxation time distribution in porous media

Microscopically the volume of a single pore in a porous media may be divided into two regions; surface area and bulk volume (Figure 1).

Figure 1: Nuclear spin relaxation properties in a simplified pore are divided into bulk volume
V
{\displaystyle V}
and pore surface area
S
{\displaystyle S}
. Nuclear spin relaxation in porous media.png
Figure 1: Nuclear spin relaxation properties in a simplified pore are divided into bulk volume and pore surface area .

The surface area is a thin layer with thickness of a few molecules close to the pore wall surface. The bulk volume is the remaining part of the pore volume and usually dominates the overall pore volume. With respect to NMR excitations of nuclear states for hydrogen-containing molecules in these regions, different relaxation times for the induced excited energy states are expected. The relaxation time is significantly shorter for a molecule in the surface area, compared to a molecule in the bulk volume. This is an effect of paramagnetic centres in the pore wall surface that causes the relaxation time to be faster. The inverse of the relaxation time , is expressed by contributions from the bulk volume , the surface area and the self-diffusion : [2]

with

where is the thickness of the surface area, is the surface area, is the pore volume, is the relaxation time in the bulk volume, is the relaxation time for the surface, is the gyromagnetic ratio, is the magnetic field gradient (assumed to be constant), is the time between echoes and is the self-diffusion coefficient of the fluid. The surface relaxation can be assumed as uniform or non-uniform. [3]

The NMR signal intensity in the distribution plot reflected by the measured amplitude of the NMR signal is proportional to the total amount of hydrogen nuclei, while the relaxation time depends on the interaction between the nuclear spins and the surroundings. In a characteristic pore containing for an example, water, the bulk water exhibits a single exponential decay. The water close to the pore wall surface exhibits faster relaxation time for this characteristic pore size.

NMR permeability correlations

NMR techniques are typically used to predict permeability for fluid typing and to obtain formation porosity, which is independent of mineralogy. The former application uses a surface-relaxation mechanism to relate measured relaxation spectra with surface-to-volume ratios of pores, and the latter is used to estimate permeability. The common approach is based on the model proposed by Brownstein and Tarr. [4] They have shown that, in the fast diffusion limit, given by the expression:

where is the surface relaxivity of pore wall material, is the radius of the spherical pore and is the bulk diffusivity. The connection between NMR relaxation measurements and petrophysical parameters such as permeability stems from the strong effect that the rock surface has on promoting magnetic relaxation. For a single pore, the magnetic decay as a function of time is described by a single exponential:

where is the initial magnetization and the transverse relaxation time is given by:

is the surface-to-volume ratio of the pore, is bulk relaxation time of the fluid that fills the pore space, and is the surface relaxation strength. For small pores or large , the bulk relaxation time is small and the equation can be approximated by:

Real rocks contain an assembly of interconnected pores of different sizes. The pores are connected through small and narrow pore throats (i.e. links) that restrict interpore diffusion. If interpore diffusion is negligible, each pore can be considered to be distinct and the magnetization within individual pores decays independently of the magnetization in neighbouring pores. The decay can thus be described as:

where is the volume fraction of pores of size that decays with relaxation time . The multi-exponential representation corresponds to a division of the pore space into main groups based on (surface-to-volume ratio) values. Due to the pore size variations, a non-linear optimization algorithm with multi-exponential terms is used to fit experimental data. [5] Usually, a weighted geometric mean, , of the relaxation times is used for permeability correlations:

is thus related to an average or pore size. Commonly used NMR permeability correlations as proposed by Dunn et al. are of the form: [6]

where is the porosity of the rock. The exponents and are usually taken as four and two, respectively. Correlations of this form can be rationalized from the Kozeny–Carman equation:

by assuming that the tortuosity is proportional to . However, it is well known that tortuosity is not only a function of porosity. It also depends on the formation factor . The formation factor can be obtained from resistivity logs and is usually readily available. This has given rise to permeability correlations of the form:

Standard values for the exponents and , respectively. Intuitively, correlations of this form are a better model since it incorporates tortuosity information through .

The value of the surface relaxation strength affects strongly the NMR signal decay rate and hence the estimated permeability. Surface relaxivity data are difficult to measure, and most NMR permeability correlations assume a constant . However, for heterogeneous reservoir rocks with different mineralogy, is certainly not constant and surface relaxivity has been reported to increase with higher fractions of microporosity. [7] If surface relaxivity data are available it can be included in the NMR permeability correlation as

relaxation

For fully brine saturated porous media, three different mechanisms contribute to the relaxation: bulk fluid relaxation, surface relaxation, and relaxation due to gradients in the magnetic field. In the absence of magnetic field gradients, the equations describing the relaxation are: [8]

on S

with the initial condition

and

where is the self-diffusion coefficient. The governing diffusion equation can be solved by a 3D random walk algorithm. Initially, the walkers are launched at random positions in the pore space. At each time step, , they advance from their current position, , to a new position, , by taking steps of fixed length in a randomly chosen direction. The time step is given by:

The new position is given by

The angles and represent the randomly selected direction for each random walker in spherical coordinates. It can be noted that must be distributed uniformly in the range (0,). If a walker encounters a pore-solid interface, it is killed with a finite probability . The killing probability is related to the surface relaxation strength by: [9]

If the walker survives, it simply bounces off the interface and its position does not change. At each time step, the fraction of the initial walkers that are still alive is recorded. Since the walkers move with equal probability in all directions, the above algorithm is valid as long as there is no magnetic gradient in the system.

When protons are diffusing, the sequence of spin echo amplitudes is affected by inhomogeneities in the permanent magnetic field. This results in an additional decay of the spin echo amplitudes that depends on the echo spacing . In the simple case of a uniform spatial gradient , the additional decay can be expressed as a multiplicative factor:

where is the ratio of the Larmor frequency to the magnetic field intensity. The total magnetization amplitude as a function of time is then given as:

NMR as a tool to measure wettability

The wettability conditions in a porous media containing two or more immiscible fluid phases determine the microscopic fluid distribution in the pore network. Nuclear magnetic resonance measurements are sensitive to wettability because of the strong effect that the solid surface has on promoting magnetic relaxation of the saturating fluid. The idea of using NMR as a tool to measure wettability was presented by Brown and Fatt in 1956. [10] The magnitude of this effect depends upon the wettability characteristics of the solid with respect to the liquid in contact with the surface. [11] Their theory is based on the hypothesis that molecular movements are slower in the bulk liquid than at the solid-liquid interface. In this solid-liquid interface the diffusion coefficient is reduced, which correspond to a zone of higher viscosity. In this higher viscosity zone, the magnetically aligned protons can more easily transfer their energy to their surroundings. The magnitude of this effect depends upon the wettability characteristics of the solid with respect to the liquid in contact with the surface.

NMR Cryoporometry for measuring pore size distributions

NMR Cryoporometry (NMRC) is a recent technique for measuring total porosity and pore size distributions. It makes use of the Gibbs-Thomson effect  : small crystals of a liquid in the pores melt at a lower temperature than the bulk liquid : The melting point depression is inversely proportional to the pore size. The technique is closely related to that of the use of gas adsorption to measure pore sizes (Kelvin equation). Both techniques are particular cases of the Gibbs Equations (Josiah Willard Gibbs): the Kelvin Equation is the constant temperature case, and the Gibbs-Thomson Equation is the constant pressure case. [12]

To make a Cryoporometry measurement, a liquid is imbibed into the porous sample, the sample cooled until all the liquid is frozen, and then warmed slowly while measuring the quantity of the liquid that has melted. Thus it is similar to DSC thermoporosimetry, but has higher resolution, as the signal detection does not rely on transient heat flows, and the measurement can be made arbitrarily slowly. It is suitable for measuring pore diameters in the range 2 nm–2 μm.

Nuclear Magnetic Resonance (NMR) may be used as a convenient method of measuring the quantity of liquid that has melted, as a function of temperature, making use of the fact that the relaxation time in a frozen material is usually much shorter than that in a mobile liquid. The technique was developed at the University of Kent in the UK. [13] It is also possible to adapt the basic NMRC experiment to provide structural resolution in spatially dependent pore size distributions, [14] or to provide behavioural information about the confined liquid. [15]

See also

Related Research Articles

Brownian motion Random motion of particles suspended in a fluid

Brownian motion or pedesis is the random motion of particles suspended in a fluid resulting from their collision with the fast-moving molecules in the fluid.

Ficks laws of diffusion mathematical descriptions of molecular diffusion

Fick's laws of diffusion describe diffusion and were derived by Adolf Fick in 1856. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

The Grashof number (Gr) is a dimensionless number in fluid dynamics and heat transfer which approximates the ratio of the buoyancy to viscous force acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number. It's believed to be named after Franz Grashof. Though this grouping of terms had already been in use, it wasn't named until around 1921, 28 years after Franz Grashof's death. It's not very clear why the grouping was named after him.

In fluid mechanics, the Rayleigh number (Ra) for a fluid is a dimensionless number associated with buoyancy-driven flow, also known as free or natural convection. It characterises the fluid's flow regime: a value in a certain lower range denotes laminar flow; a value in a higher range, turbulent flow. Below a certain critical value, there is no fluid motion and heat transfer is by conduction rather than convection.

Skin effect

Skin effect is the tendency of an alternating electric current (AC) to become distributed within a conductor such that the current density is largest near the surface of the conductor, and decreases with greater depths in the conductor. The electric current flows mainly at the "skin" of the conductor, between the outer surface and a level called the skin depth. The skin effect causes the effective resistance of the conductor to increase at higher frequencies where the skin depth is smaller, thus reducing the effective cross-section of the conductor. The skin effect is due to opposing eddy currents induced by the changing magnetic field resulting from the alternating current. At 60 Hz in copper, the skin depth is about 8.5 mm. At high frequencies the skin depth becomes much smaller. Increased AC resistance due to the skin effect can be mitigated by using specially woven litz wire. Because the interior of a large conductor carries so little of the current, tubular conductors such as pipe can be used to save weight and cost.

The nuclear Overhauser effect (NOE) is the transfer of nuclear spin polarization from one population of spin-active nuclei to another via cross-relaxation. A phenomenological definition of the NOE in nuclear magnetic resonance spectroscopy (NMR) is the change in the integrated intensity of one NMR resonance that occurs when another is saturated by irradiation with an RF field. The change in resonance intensity of a nucleus is a consequence of the nucleus being close in space to those directly affected by the RF perturbation.

Permeability in fluid mechanics and the Earth sciences is a measure of the ability of a porous material to allow fluids to pass through it.

Darcy's law is an equation that describes the flow of a fluid through a porous medium. The law was formulated by Henry Darcy based on results of experiments on the flow of water through beds of sand, forming the basis of hydrogeology, a branch of earth sciences.

In physics, Washburn's equation describes capillary flow in a bundle of parallel cylindrical tubes; it is extended with some issues also to imbibition into porous materials. The equation is named after Edward Wight Washburn; also known as Lucas–Washburn equation, considering that Richard Lucas wrote a similar paper three years earlier, or the Bell-Cameron-Lucas-Washburn equation, considering J.M. Bell and F.K. Cameron's discovery of the form of the equation in 1906.

In the physical sciences, relaxation usually means the return of a perturbed system into equilibrium. Each relaxation process can be categorized by a relaxation time τ. The simplest theoretical description of relaxation as function of time t is an exponential law exp(-t/τ).

Hydraulic conductivity, symbolically represented as , is a property of vascular plants, soils and rocks, that describes the ease with which a fluid can move through pore spaces or fractures. It depends on the intrinsic permeability of the material, the degree of saturation, and on the density and viscosity of the fluid. Saturated hydraulic conductivity, Ksat, describes water movement through saturated media. By definition, hydraulic conductivity is the ratio of velocity to hydraulic gradient indicating permeability of porous media.

The Tolman length measures the extent by which the surface tension of a small liquid drop deviates from its planar value. It is conveniently defined in terms of an expansion in , with the equimolar radius of the liquid drop, of the pressure difference across the droplet's surface:

Nonlinear acoustics (NLA) is a branch of physics and acoustics dealing with sound waves of sufficiently large amplitudes. Large amplitudes require using full systems of governing equations of fluid dynamics and elasticity. These equations are generally nonlinear, and their traditional linearization is no longer possible. The solutions of these equations show that, due to the effects of nonlinearity, sound waves are being distorted as they travel.

Diffusivity, mass diffusivity or diffusion coefficient is a proportionality constant between the molar flux due to molecular diffusion and the gradient in the concentration of the species. Diffusivity is encountered in Fick's law and numerous other equations of physical chemistry.

In physics and engineering, mass flux is the rate of mass flow per unit area, perfectly overlapping with the momentum density, the momentum per unit volume. The common symbols are j, J, q, Q, φ, or Φ, sometimes with subscript m to indicate mass is the flowing quantity. Its SI units are kg s−1 m−2. Mass flux can also refer to an alternate form of flux in Fick's law that includes the molecular mass, or in Darcy's law that includes the mass density.

Porosity or void fraction is a measure of the void spaces in a material, and is a fraction of the volume of voids over the total volume, between 0 and 1, or as a percentage between 0% and 100%. Strictly speaking, some tests measure the "accessible void", the total amount of void space accessible from the surface. There are many ways to test porosity in a substance or part, such as industrial CT scanning. The term porosity is used in multiple fields including pharmaceutics, ceramics, metallurgy, materials, manufacturing, hydrology, earth sciences, soil mechanics and engineering.

Diffusion Net movement of molecules or atoms from a region of high concentration (or high chemical potential) to a region of low concentration (or low chemical potential)

Diffusion is net movement of anything from a region of higher concentration to a region of lower concentration. Diffusion is driven by a gradient in concentration. For example, if you spray perfume at one end of a room eventually the gas particles will be all over the room

The Gibbs–Thomson effect, in common physics usage, refers to variations in vapor pressure or chemical potential across a curved surface or interface. The existence of a positive interfacial energy will increase the energy required to form small particles with high curvature, and these particles will exhibit an increased vapor pressure. See Ostwald–Freundlich equation. More specifically, the Gibbs–Thomson effect refers to the observation that small crystals are in equilibrium with their liquid melt at a lower temperature than large crystals. In cases of confined geometry, such as liquids contained within porous media, this leads to a depression in the freezing point / melting point that is inversely proportional to the pore size, as given by the Gibbs–Thomson equation.

The Strange–Rahman–Smith equation is used in the cryoporometry method of measuring porosity. NMR cryoporometry is a recent technique for measuring total porosity and pore size distributions. NMRC is based on two equations: the Gibbs–Thomson equation, which maps the melting point depression to pore size, and the Strange–Rahman–Smith equation, which maps the melted signal amplitude at a particular temperature to pore volume.

In fluid mechanics, fluid flow through porous media is the manner in which fluids behave when flowing through a porous medium, for example sponge or wood, or when filtering water using sand or another porous material. As commonly observed, some fluid flows through the media while some mass of the fluid is stored in the pores present in the media.

References

  1. Allen, S.G.; Stephenson, P.C.L.; Strange, J.H. (1997), "Morphology of porous media studied by nuclear magnetic resonance", Journal of Chemical Physics , 106 (18): 7802, Bibcode:1997JChPh.106.7802A, doi:10.1063/1.473780
  2. Brownstein, K.R.; Tarr, C.E. (1977), "Spin-lattice relaxation in a system governed by diffusion", Journal of Magnetic Resonance , 26: 17–24, doi:10.1016/0022-2364(77)90230-X
  3. Valfouskaya, A.; Adler, P.M.; Thovert, J.F.; Fleury, M. (2005), "Nuclear magnetic resonance diffusion with surface relaxation in porous media", Journal of Colloid and Interface Science , 295 (1): 188–201, Bibcode:2006JCIS..295..188V, doi:10.1016/j.jcis.2005.08.021, PMID   16168421
  4. Brownstein, K.R.; Tarr, C.E. (1979), "Importance of classical diffusion in NMR studies of water in biological cells", Physical Review A , 19 (6): 2446, Bibcode:1979PhRvA..19.2446B, doi:10.1103/PhysRevA.19.2446
  5. Howard, J.J.; Spinler, E.A. (1995), "Nuclear magnetic resonance measurements of wettability and fluid saturations in chalk", SPE Advanced Technology Series , 3: 60–65, doi:10.2118/26471-PA
  6. Dunn, K.J.; LaTorraca, D.; Bergmann, D.J. (1999), "Permeability relation with other petrophysical parameters for periodic porous media", Geophysics , 64 (2): 470, Bibcode:1999Geop...64..470D, doi:10.1190/1.1444552
  7. Kenyon, W.E. (1992), "Nuclear magnetic resonance as a petrophysical measurement", Nuclear Geophysics , 6 (2): 153
  8. Cohen, M.H.; Mendelson, K.S. (1982), "Nuclear magnetic relaxation and the internal geometry of sedimentary rocks", Journal of Applied Physics , 53 (2): 1127, Bibcode:1982JAP....53.1127C, doi:10.1063/1.330526
  9. Bergmann, D.J.; Dunn, K.J.; Schwartz, L.M.; Mitra, P.P. (1995), "Self-diffusion in periodic porous medium: A comparison of different approaches", Physical Review E , 51 (4): 3393, Bibcode:1995PhRvE..51.3393B, doi:10.1103/PhysRevE.51.3393
  10. Brown, R.J.S.; Fatt, I. (1956), "Measurements of Fractional Wettability of Oilfield Rocks by the Nuclear Magnetic Relaxation Method", Transactions of the American Institute of Mining, Metallurgical and Petroleum Engineers , 207: 262
  11. Howard, J.J. (1998), "Quantitative estimates of porous media wettability from proton NMR", Magnetic Resonance Imaging, 16 (5–6): 529–33, doi:10.1016/S0730-725X(98)00060-5, PMID   9803903
  12. Mitchell, J.; Webber, J. B. W.; Strange, J. H. (2008), "Nuclear Magnetic Resonance Cryoporometry" (PDF), Physics Reports , 461 (1): 1–36, Bibcode:2008PhR...461....1M, doi:10.1016/j.physrep.2008.02.001
  13. Strange, J.H.; Rahman, M.; Smith, E.G. (1993), "Characterization of Porous Solids by NMR", Physical Review Letters , 71 (21): 3589–3591, Bibcode:1993PhRvL..71.3589S, doi:10.1103/PhysRevLett.71.3589, PMID   10055015
  14. Strange, J.H.; Webber, J.B.W. (1997), "Spatially resolved pore size distributions by NMR" (PDF), Measurement Science and Technology , 8 (5): 555–561, Bibcode:1997MeScT...8..555S, doi:10.1088/0957-0233/8/5/015
  15. Alnaimi, S.M.; Mitchell, J.; Strange, J.H.; Webber, J.B.W. (2004), "Binary liquid mixtures in porous solids" (PDF), Journal of Chemical Physics , 120 (5): 2075–2077, Bibcode:2004JChPh.120.2075A, doi:10.1063/1.1643730, PMID   15268344