Product operator formalism

Last updated

In NMR spectroscopy, the product operator formalism is a method used to determine the outcome of pulse sequences in a rigorous but straightforward way. With this method it is possible to predict how the bulk magnetization evolves with time under the action of pulses applied in different directions. It is a net improvement from the semi-classical vector model which is not able to predict many of the results in NMR spectroscopy and is a simplification of the complete density matrix formalism.

Contents

In this model, for a single spin, four base operators exist: , , and which represent respectively polarization (population difference between the two spin states), single quantum coherence (magnetization on the xy plane) and the unit operator. Many other, non-classical operators exist for coupled systems. Using this approach, the evolution of the magnetization under free precession is represented by and corresponds to a rotation about the z-axis with a phase angle proportional to the chemical shift of the spin in question:

Pulses about the x and y axis can be represented by and respectively; these allow to interconvert the magnetization between planes and ultimately to observe it at the end of a sequence. Since every spin will evolve differently depending on its shift, with this formalism it is possible to calculate exactly where the magnetization will end up and hence devise pulse sequences to measure the desired signal while excluding others.

The product operator formalism is particularly useful in describing experiments in two-dimensions like COSY, HSQC and HMBC.

Motivation for sets of spin-1/2 particles

Throughout this section, the reduced Planck constant for convenience.

The product operator formalism is usually applied to sets of spin-1/2 particles, since the fact that the individual operators satisfy , where is the identity operator, makes the commutation relations of product operators particularly simple. In principle the formalism could be extended to higher spins, but in practice the general irreducible spherical tensor treatment is more often used. As such, we consider only the spin-1/2 case below.

The main idea of the formalism is to make it easier to follow the system density operator , which evolves under a Hamiltonian according to the Liouville-von Neumann equation as

For a time-independent Hamiltonian, the density operator inherits its solutions from the Schrödinger time-evolution operator as

Density operator-state duality

Suppose a single spin-1/2 is in the state , which is an eigenstate of the z-spin operator , that is . Similarly . Making use of the expansion of a Hermitian operator in terms of projections onto its eigenkets with eigenvalues as , the associated density operator is

where is the identity operator. Similarly, the density operator for the state is

Since the spin operators are all traceless and the expectation value of an operator for a system with density operator is , the terms proportional to the unit operator do not affect the expectations of the spin operators. Additionally those parts do not evolve in time, since they trivially commute with the Hamiltonian. Therefore those terms can be ignored, and the state corresponds to a density operator , while the state corresponds to a density operator . In exactly the same manner, polarisation along the positive x-axis, that is a state , corresponds to a density operator . This idea extends naturally to multiple spins, where the states and operators are direct products of single-spin states and operators. Hence operator terms in the density operator have a direct duality with states.

In the case of two spins , the terms in the density operator (ignoring the identity on its own) can be interpreted as representing

where eg is a shorthand for the Kronecker product , where is the identity operator on the spin, and similarly is a shorthand for .

The factors of two in the 'true' two-spin operators are to allow for convenient commutation relations in this specific spin-1/2 case - see below. Note also that we could instead choose to expand the density operator in the basis etc, where the transverse operators have been replaced with raising and lowering operators. With quadrature detection, the observable associated with an individual spin is effectively the non-Hermitian , so this is sometimes more convenient.

Evolution of the density operator

Consider operators that obey the cyclic commutation relations

In fact only the first two relations are necessary for the following derivation, but since we are usually working with operators associated with Cartesian directions, such as the individual angular momentum operators, the third commutator follows by a symmetry argument.

Introduce also the commutation superoperator of an operator (in our case, this is more formally related to the adjoint representation of the Lie algebra whose elements are ), which acts as

In particular, for the cyclic operators, we have

and consequently for integer

An identity for two operators is

which can be derived by putting where is a scalar parameter, differentiating both sides with respect to , and noting that both sides satisfy the same differential equation in that parameter, with the same initial condition at . In particular, for some scalar parameter , we have

where the final equality follows from recognising the Taylor series for sine and cosine. Now suppose that the density operator at time zero is , and it is allowed to freely evolve under the Hamiltonian where is some scalar. Using the results above, the density operator at some later time will be given by

The interpretation of this is that although nuclear spin angular momentum itself is not connected to rotations in three-dimensional space in the same way that angular momentum is, the evolution of the density operator can be viewed as rotations in an abstract space, in which the operators are the generators of rotations about the axes. An example of such a set of generators is just the spin operators themselves.

We now also introduce the 'arrow notation' typically used in NMR, which writes the general evolution given above as the shorthand

.

With more specific reference to the radiofrequency pulses applied during NMR experiments, a hard pulse with tip angle around a direction is written as above the arrow and corresponds to taking as the rotation generator in Equation 1 . When there is no ambiguity, the arrow label may be omitted, or be eg text instead.

Note that a more complicated calculation has now been reduced to a simpler procedure that requires no knowledge of the underlying quantum mechanics, especially since the subspaces of cyclic operators can be tabulated in advance.

Examples

The 180°-refocussing pulse

The Hamiltonian for a single spin evolving under a chemical shift of angular frequency is

which means that in an ensemble of many such spins with slightly different chemical shifts, there is a dephasing of the magnetisation in the - plane. Consider the pulse sequence

where is a time interval. Starting in an equilibrium state with all the polarisation along the -axis, the evolution of an individual spin in the ensemble is

Hence this sequence refocuses the transverse magnetisation produced by the first pulse, independent of the value of the chemical shift.

As an indication of the utility of the formalism, suppose instead that we tried to reach the same result using states only and therefore the Schrödinger time evolution operators. This amounts to trying to simplify the unitary propagator taking the initial state to the final state as , where explicitly

Essentially we want to find the propagator in the form , that is as a single exponential of a combination of operators, because that gives the effective Hamiltonian acting during the sequence. Since the arguments of the exponentials in the original form of the propagator do not commute, this amounts to solving a specific example of the Baker–Campbell–Hausdorff (BCH) problem. In this relatively simple case we can solve the BCH problem using the fact that for unitary operator , operator and function , as well as the mathematical similarity of the spin operators with the physical rotation generators, which allow us to write

Hence and only the effect of the 180° pulse remains, which agrees with the product operator treatment. For larger sequences of pulses this state treatment quickly becomes even more unwieldy, unless more advanced methods such as exact effective Hamiltonian theory (which gives closed-form expressions for the entangled propagators via the Cayley–Hamilton theorem and eigendecompositions) are used.

The amplitude of a Hahn echo in an inhomogeneous magnetic field

As an extension of the refocussing pulse treated above, consider a set of two pulses with arbitrary flip angles and , that is sequence

where again is a time interval. Liberally dropping irrelevant terms, the evolution for a single spin with offset up to just after the second pulse is

Now consider an ensemble of spins in a magnetic field that is sufficiently inhomogeneous to completely dephase the spins in the interval between the pulses. After the second pulse, we can decompose the remaining terms into a sum of two spin populations differing only in the sign of the term, in the sense that for an individual spin we have

where we used the identities and .

It is the spins in the new population that has been generated by the second pulse, namely the one with , that will lead to the formation of an echo after evolution for the next interval. Therefore, remembering to include the introduced by the first pulse, the amplitude of the resulting Hahn echo relative to that produced by an ideal 90°—180° refocussing pulse sequence is roughly

Note that this is not an exact result, because it considers only the refocussing of polarisation that was transverse immediately before the second pulse. In reality there will be further transverse components originating from the tipping of the longitudinal magnetisation that remained after the first pulse. However, for many tip angles, this is a good rule of thumb.

To instead arrive at this result using the state formalism, we would have had to non-trivially evaluate the rotation propagator as

and then evaluate a transition probability by considering the result of applying this to a state representing polarisation in the transverse plane.

DEPT (Distortionless Enhancement by Polarisation Transfer)

DEPT (Distortionless Enhancement by Polarisation Transfer) is a pulse sequence used to distinguish between the multiplicity of hydrogen bonded to carbon, that is it can separate C, CH, CH2 and CH3 groups. It does this by exploiting the heteronuclear carbon-hydrogen -coupling and varying the tip angle of the final pulse in the sequence. The basic pulse sequence is shown below.

DEPT basic pulse sequence.svg

Under the weak coupling assumption, the chemical shift terms commute with the -coupling term in the Hamiltonian. Hence we can ignore the refocussed chemical shift (see § The 180°-refocussing pulse) in the two intervals containing -pulses, namely and , and additionally refrain from evaluating the chemical shift evolution in the last period . The pulse separation time is adjusted to the coupling strength (with associated Hamiltonian coefficient ) such that it satisfies

,

because then the first term in the evolved density operator in Equation 2 vanishes under the pure coupling evolution between the pulses.

CH

Label the hydrogen spin as , and the carbon spin by . For illustrative purposes, we assume that the equilibrium state only has polarisation on the -spin (in reality, there will also be polarisation on the spin, with the relative populations determined by the thermal Boltzmann factors). The -coupling Hamiltonian is

which gives the following evolution

The non-trivial commutators used to identify the cyclic subspace for are

and consequently the next cyclic rotation

where we used the 'mixed-product identity' , which relates the matrix and Kronecker products for compatible dimensions of , and also the fact that since the two eigenvalues of any of the spin-1/2 operators are , any of their squares are given by by the Cayley–Hamilton theorem.

Note also that the term is invariant under the -coupling evolution. That is that the term commutes with the Hamiltonian, and in this case, that can be manually confirmed by evaluating the commutator using the matrix representations of the spin operators.

CH2

Now label the two hydrogen spins as and the carbon spin by . The -coupling Hamiltonian is now

which gives the following evolution

where 'others' denotes various terms that can safely be ignored because they will not evolve into observable transverse polarisation on the target spin . The required cyclic commutators for dealing with the -coupling evolution are the following three sets (and their versions if needed)

CH3

A similar (but more lengthy) treatment gives the final observable term as .

APT (Attached Proton Test)

Refer to § DEPT (Distortionless Enhancement by Polarisation Transfer) for the notation used in this example.

APT is similar to DEPT in that it detects carbon multiplicity. However, it has additional degeneracies: it gives identical positive signals for C and CH2, and identical negative signals for CH and CH3. One variation on the basic pulse sequence is shown below.

APT basic pulse sequence.svg

The key observation is that since we can again ignore the refocussed chemical shift, the only relevant dynamics occur in the interval with no hydrogen decoupling, where we can consider solely the -coupling. By using an interval twice as long as in the DEPT case, we ensure that a density operator of at the start of the interval just has its sign inverted following the coupling (since this corresponds to in the general treatment, and ). The Hamiltonians for the couplings to each of the separate neighbouring hydrogen atoms commute, so the overall effect is to multiply by a factor . This motivates the alternating sign of the signal mentioned above.

Related Research Articles

In electrodynamics, linear polarization or plane polarization of electromagnetic radiation is a confinement of the electric field vector or magnetic field vector to a given plane along the direction of propagation. The term linear polarization was coined by Augustin-Jean Fresnel in 1822. See polarization and plane of polarization for more information.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

<span class="mw-page-title-main">Spherical harmonics</span> Special mathematical functions defined on the surface of a sphere

In mathematics and physical science, spherical harmonics are special functions defined on the surface of a sphere. They are often employed in solving partial differential equations in many scientific fields. The table of spherical harmonics contains a list of common spherical harmonics.

An infinitesimal rotation matrix or differential rotation matrix is a matrix representing an infinitely small rotation.

<span class="mw-page-title-main">Rabi cycle</span> Quantum mechanical phenomenon

In physics, the Rabi cycle is the cyclic behaviour of a two-level quantum system in the presence of an oscillatory driving field. A great variety of physical processes belonging to the areas of quantum computing, condensed matter, atomic and molecular physics, and nuclear and particle physics can be conveniently studied in terms of two-level quantum mechanical systems, and exhibit Rabi flopping when coupled to an optical driving field. The effect is important in quantum optics, magnetic resonance and quantum computing, and is named after Isidor Isaac Rabi.

In quantum physics, the scattering amplitude is the probability amplitude of the outgoing spherical wave relative to the incoming plane wave in a stationary-state scattering process. At large distances from the centrally symmetric scattering center, the plane wave is described by the wavefunction

<span class="mw-page-title-main">Rotating reference frame</span> Concept in classical mechanics

A rotating frame of reference is a special case of a non-inertial reference frame that is rotating relative to an inertial reference frame. An everyday example of a rotating reference frame is the surface of the Earth.

<span class="mw-page-title-main">Etendue</span> Measure of the "spread" of light in an optical system

Etendue or étendue is a property of light in an optical system, which characterizes how "spread out" the light is in area and angle. It corresponds to the beam parameter product (BPP) in Gaussian beam optics. Other names for etendue include acceptance, throughput, light grasp, light-gathering power, optical extent, and the AΩ product. Throughput and AΩ product are especially used in radiometry and radiative transfer where it is related to the view factor. It is a central concept in nonimaging optics.

In special functions, a topic in mathematics, spin-weighted spherical harmonics are generalizations of the standard spherical harmonics and—like the usual spherical harmonics—are functions on the sphere. Unlike ordinary spherical harmonics, the spin-weighted harmonics are U(1) gauge fields rather than scalar fields: mathematically, they take values in a complex line bundle. The spin-weighted harmonics are organized by degree l, just like ordinary spherical harmonics, but have an additional spin weights that reflects the additional U(1) symmetry. A special basis of harmonics can be derived from the Laplace spherical harmonics Ylm, and are typically denoted by sYlm, where l and m are the usual parameters familiar from the standard Laplace spherical harmonics. In this special basis, the spin-weighted spherical harmonics appear as actual functions, because the choice of a polar axis fixes the U(1) gauge ambiguity. The spin-weighted spherical harmonics can be obtained from the standard spherical harmonics by application of spin raising and lowering operators. In particular, the spin-weighted spherical harmonics of spin weight s = 0 are simply the standard spherical harmonics:

Sinusoidal plane-wave solutions are particular solutions to the electromagnetic wave equation.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

In geometry, various formalisms exist to express a rotation in three dimensions as a mathematical transformation. In physics, this concept is applied to classical mechanics where rotational kinematics is the science of quantitative description of a purely rotational motion. The orientation of an object at a given instant is described with the same tools, as it is defined as an imaginary rotation from a reference placement in space, rather than an actually observed rotation from a previous placement in space.

The Wigner D-matrix is a unitary matrix in an irreducible representation of the groups SU(2) and SO(3). It was introduced in 1927 by Eugene Wigner, and plays a fundamental role in the quantum mechanical theory of angular momentum. The complex conjugate of the D-matrix is an eigenfunction of the Hamiltonian of spherical and symmetric rigid rotors. The letter D stands for Darstellung, which means "representation" in German.

<span class="mw-page-title-main">Axis–angle representation</span> Parameterization of a rotation into a unit vector and angle

In mathematics, the axis–angle representation parameterizes a rotation in a three-dimensional Euclidean space by two quantities: a unit vector e indicating the direction (geometry) of an axis of rotation, and an angle of rotation θ describing the magnitude and sense of the rotation about the axis. Only two numbers, not three, are needed to define the direction of a unit vector e rooted at the origin because the magnitude of e is constrained. For example, the elevation and azimuth angles of e suffice to locate it in any particular Cartesian coordinate frame.

<span class="mw-page-title-main">Diffraction from slits</span>

Diffraction processes affecting waves are amenable to quantitative description and analysis. Such treatments are applied to a wave passing through one or more slits whose width is specified as a proportion of the wavelength. Numerical approximations may be used, including the Fresnel and Fraunhofer approximations.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

In physics, and especially scattering theory, the momentum-transfer cross section is an effective scattering cross section useful for describing the average momentum transferred from a particle when it collides with a target. Essentially, it contains all the information about a scattering process necessary for calculating average momentum transfers but ignores other details about the scattering angle.

In quantum information theory, the Wehrl entropy, named after Alfred Wehrl, is a classical entropy of a quantum-mechanical density matrix. It is a type of quasi-entropy defined for the Husimi Q representation of the phase-space quasiprobability distribution. See for a comprehensive review of basic properties of classical, quantum and Wehrl entropies, and their implications in statistical mechanics.

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

In the hyperbolic plane, as in the Euclidean plane, each point can be uniquely identified by two real numbers. Several qualitatively different ways of coordinatizing the plane in hyperbolic geometry are used.

References