You are on page 1of 38

Computer Simulations in

Solid-State NMR. I. Spin


Dynamics Theory
MATTIAS EDÉN
Physical Chemistry Division, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden

ABSTRACT: This article is the first in a series of publications that discuss the basics of
the writing of computer programs for simulating solid-state NMR experiments with static
and rotating samples. The present article gives an account of the relevant NMR theory
needed for writing NMR simulation computer codes. The concept of irreducible spherical
tensors is reviewed, as is how it may be used to construct Hamiltonians that represent the
various NMR interactions. The spin dynamics that links the Hamiltonian to the time
evolution of the nuclear spins and the detection of the NMR time-domain signal is
discussed for static and rotating solids, as well as the relationship between the form of the
Hamiltonian and the resulting NMR spectrum. © 2003 Wiley Periodicals, Inc. Concepts
Magn Reson Part A 17A: 117–154, 2003

KEY WORDS: solid-state NMR; numerical simulation; static solids; magic-angle spinning;
dynamically inhomogeneous Hamiltonian; spherical tensor

INTRODUCTION study the relative importance of various parameters


affecting the NMR measurement. Such investigations
The analysis of most NMR experiments relies on may prove difficult or even impossible to carry out
numerical simulations. For example, numerical simu- experimentally.
lations are needed for extracting information about All NMR interactions have anisotropic contribu-
spin interaction parameters and molecular structure by tions (1– 6), which is clearly manifested in NMR
means of iterative fitting of calculated NMR re- experiments with solid samples. The NMR response
sponses to experimental data. Computer modeling of of the interactions depends on the orientation of the
experiments are also extensively used for testing and molecule (containing the nuclear spins under study)
optimizing new NMR methodologies: through a com- with respect to the direction of the static magnetic
puter program, it is straightforward to isolate and field. Experiments are usually performed on powders
comprising a large number of randomly oriented crys-
Received 5 October 2001; revised 22 October 2002; tallites. The resulting NMR spectrum is then a super-
accepted 20 November 2002 position of the spectra from all crystallites. This gives
Correspondence to: Dr. M. Edén; E-mail: mattias@physc.su.se powder spectra with NMR resonances spread over a
Concepts in Magnetic Resonance Part A, Vol. 17A(1) 117–154 (2003) large range of frequencies, as seen in Fig. 1(a).
Published online in Wiley InterScience (www.interscience.wiley. In order to obtain high resolution NMR spectra
com). DOI 10.1002/cmr.a.10061 from powders, experimental manipulations of the
© 2003 Wiley Periodicals, Inc. spins are often required, such as magic-angle spinning
117
118 EDÉN

cause the calculation must then incorporate powder


averaging, implying a summation of the signals from
a large number of molecular orientations.
Over the last few years, simulations of solid-state
NMR experiments have reached a high degree of
sophistication. It is possible to simulate NMR signals
by mimicking the experimental situation very closely,
using a minimum of assumptions. For example, pow-
der averaged MAS spectra from multiple-spin sys-
tems can be simulated in a few seconds or minutes on
a personal computer, taking all relevant interactions
of the spin system into account (9 –14). Even powder
averaged spectra for 2-dimensional experiments in-
volving complex pulse sequences may be calculated
within 1 h (14 –16). This is a result of steadily in-
creasing computer capabilities, combined with highly
efficient general-purpose computational methods that
have recently been introduced (10 –13, 15, 17–20).
There are simulation packages available for various
solid-state NMR simulations (21–23). In particular,
two very sophisticated simulation platforms, GAMMA
(24, 25) and SIMPSON (14, 26), have found well-
spread public usage.
The present article is the first in a series that seeks
to outline how a numerical simulation program may
be written for calculating NMR time-domain signals
or frequency-domain spectra with emphasis on so-
Figure 1 The experimental 13C spectra of a powder of
called dynamically inhomogeneous cases as defined
99%-13C2-labeled glycine, acquired at a magnetic field of by Maricq and Waugh (27). These correspond to
9.4 T. High-power proton decoupling was employed during evolution of the nuclear spins under either a time-
the acquisition. (a) The spectrum of a static sample. (b– d) independent Hamiltonian or a time-dependent Ham-
Spectra for a sample rotating at the magic angle and re- iltonian that commutes with itself at all times. This
corded at a spinning frequency equal to (b) 2.50, (c) 5.30, applies, for instance, to isolated spins and hetero-
and (d) 10.00 kHz. nuclear spin systems under MAS conditions. These
articles may also be of interest to the users of already
existing simulation software, because we discuss gen-
(MAS) and/or techniques involving high power radio eral aspects of utility when executing simulation pro-
frequency (RF) field application (1–5, 7, 8). MAS is grams.
based on the idea that because the anisotropic inter- Writing a numerical simulation program requires
actions are orientation dependent, they may be re- insight into the NMR theory underlying the descrip-
moved by rapidly rotating the sample about the axis of tion of the experiment, in particular, the form of the
the sample holder, which should be inclined at the Hamiltonian under which the spins evolve. Moreover,
“magic angle” ␪ m ⫽ arctan公2 with respect to the it is necessary to have a systematic procedure for
static magnetic field direction. MAS diminishes the constructing the spin Hamiltonian for any given ex-
effects of the anisotropic interactions; and if the sam- periment. In this series of Concepts articles, we ad-
ple rotation is fast enough, they are averaged out dress the problem using the irreducible spherical ten-
completely, as shown in Fig. 1(d). sor (IST) formalism (1–5, 28, 29). An explanation is
As a result of the time-dependent modulations, made on how the Hamiltonian may be formed as a
combined with the anisotropic nature of the spin in- product of one IST (the “spatial tensor”) that depends
teractions, numerical simulations of solid-state NMR on the orientation of the molecule with respect to the
experiments are quite demanding and generally re- magnetic field and another IST that is represented by
quire different computational techniques than those spin operators. The IST formalism is highly suited for
for isotropic liquids. Additional complications arise describing the response of NMR interactions to rota-
when simulating NMR responses from powders, be- tions of the sample, which is needed for the theoret-
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 119

ical, as well as numerical, treatment of MAS experi- MATHEMATICAL CONCEPTS AND


ments. GENERAL CONVENTIONS
A numerical simulation program for calculating
the NMR time-domain signal or frequency-domain This section reviews basic mathematical tools em-
spectrum comprises the following main parts: ployed in the formulation of spin dynamics theory.
They are used later in the implementations of the
computer codes in subsequent articles. The following
1. Initialization. For numerical simulations, it is references contain additional information on these
first necessary to provide the program with in- topics: diagonalization of operators and exponential
put data characterizing the spin system and its operators (2– 4, 6, 29, 30); Fourier transformation
interactions. Parameters describing the particu- (4, 6, 30); Euler angles, tensors, and rotations (1, 4,
lar experimental situation are also required, 5, 29).
such as the MAS frequency. Furthermore, the
desired form of output of the calculation is
needed, for example, the spectral window and Notation
frequency resolution. Simulations of complex Entities that may be represented as matrices (such as
experiments may require a large amount of in- tensors and vectors) are set in boldface letters. Vec-
put data. Once the necessary data are gathered, tors are superscripted by single-headed arrows (e.g.,
7
the spatial tensors and spin operators are con- nᠬ ) and tensors by double-headed arrows (e.g., A ). All
structed, as discussed in later sections. operators are superscripted by a circumflex (Îz ).
2. Spin dynamics calculation. This is the heart of Components of vectors and tensors are italicized
the simulation program, where typically 99% of (e.g., V z and A lm ); and, if the tensor components
the computational time is spent and finally re- themselves are operators, they are set in bold letters
sults in the calculated NMR response. Calculat- (e.g., T̂lm ). We employ two alternative notations for
ing “spin dynamics” means finding how the matrix elements: in the outline of numerical algo-
spin density operator (which specifies the state rithms, O jk denotes the element of row j and column
of an ensemble of spin systems) evolves in time k of the operator Ô, evaluated in some basis set; and
during the NMR experiment that is to be sim- we also make frequent use of the Dirac (bracket)
ulated. The nature of the spin dynamics is di- formalism (2, 3, 29). For a given basis, the relation-
rectly reflected in the NMR time-domain signal ship between the two notations is O jk ⬅ 具j兩Ô兩k典.
and spectrum. This article outlines a spin dy- An object A represented in a coordinate system
namics theory, which is examined in detail for (frame) F is denoted within brackets, and the frame
label is superscript: [A] F .
spin systems evolving under dynamically inho-
We also make a distinction between points in time,
mogeneous Hamiltonians. The relationship be-
denoted t, and time intervals (i.e., differences between
tween the NMR spectrum and the particular
two points in time), denoted ␶ (31).
type of Hamiltonian (time independent as in the
case of a static solid or time periodic as in the
case of MAS) is also examined. Diagonalization of Operators
3. Postprocessing. The final part of the simulation
An Hermitian operator (2, 3, 6, 29) is invariant to
program may involve additional processing of
taking transpose and complex conjugate (together rep-
the calculated NMR signal, such as applying
resented by the adjoint operation †)
broadening to the spectral peaks.

 ⫽ † [1]
The numerical implementations of the spin dynam-
ics calculations and postprocessing of the NMR sig- Any Hermitian operator  (e.g., the Hamiltonian Ĥ)
nals will be discussed in the following article. We may be diagonalized, that is, brought into a represen-
stress that this article only deals with calculation of tation such that all its matrix elements are zero, except
the NMR response from a single molecular orienta- those on the diagonal. We denote this representation
tion (i.e., a single crystal). Simulations of time signals Âdiag. The particular basis where the operator is diag-
and spectra of powders require powder averaging. onal is called the eigenbasis of the operator. For an
This stage of the simulation will be included in a operator of dimensions ᏺ ⫻ ᏺ, the eigenbasis is
following publication. spanned by ᏺ orthonormal eigenstates (eigenvectors)
120 EDÉN

of Â: {兩1典, 兩2典, . . . , 兩ᏺ典}. The orthonormality condi- Multiplying Eq. [7] from the left with 具k兩 and from the
tion means that the scalar product between a state with right with 兩k典 gives
itself is one and that between any two different states
is zero:

1 if j ⫽ k
具j兩k典 ⫽ ␦共 j, k兲 ⫽ 0 if j ⫽ k再 [2]
具k兩Â diag兩k典 ⫽ 具k兩 冉冘 冊

j⫽1
aj 兩j典具j兩 兩k [8]

The Dirac formalism allows us to interpret each term


where ␦ represents the Kronecker ␦ function (32). of the right-hand side either as a j 具k兩 䡠 兩j典具j兩 䡠 兩k典 or as
Each of the eigenstates of  obeys the following a product of two scalar products a j 具k兩j典 䡠 具j兩k典 (29).
so-called eigenequation (2– 4, 6, 29): Using the latter interpretation gives

Â兩j典 ⫽ aj 兩j典 [3]


冘 a 具k兩j典 䡠 具j兩k典

具k兩Â diag兩k典 ⫽ j [9]


It means that if the operator acts on one of its eigen- j⫽1
states 兩j典, the outcome is the same state, multiplied by
a number a j , called the eigenvalue of  that corre-

冘 a ␦共 j, k兲

sponds to 兩j典.
To diagonalize Â, the operator that effects the ⫽ j [10]
transformation  3 Âdiag is needed. This operator is j⫽1

denoted X̂ and its matrix representation has the nor-


malized eigenstates of  as columns. In addition, X̂ is From the orthonormality of the eigenstates (Eq. [2]),
a unitary operator (29), which is defined as having the all terms in the sum vanish, except for that with j ⫽
adjoint operator as its inverse (i.e., X̂† ⫽ X̂⫺1), im- k, and we get 具k兩Âdiag兩k典 ⫽ a k as expected.
plying that Equation [5] may be used to transform an arbitrary
operator B̂ into the eigenbasis of  according to (29)
X̂X̂ † ⫽ X̂† X̂ ⫽ 1̂ [4]
B̂ 3 X̂† B̂X̂ [11]
where 1̂ is the unity operator, for which all matrix
elements are zero, except those on the diagonal: Note, however, that unless  and B̂ share the same
具j兩1̂兩k典 ⫽ ␦( j, k). The diagonal matrix Âdiag may be eigenbasis, the operator X̂†B̂X̂ is not represented by a
obtained from the following operator “sandwich”: diagonal matrix (29).

 diag ⫽ X̂† ÂX̂ [5]


Exponential Operators

冢 冣
a1 0 0 ··· 0 In the following we frequently need to form the com-
0 a2 0 ··· 0 plex exponential exp{iÂ} of an Hermitian operator
⫽ 0 0 a3 ··· 0 [6] Â. Just as the exponential of a real or complex number
· · · ·· ·
· · · · ·· c may be calculated from a Taylor series
· · ·
0 0 0 · · · aᏺ
1 2 1 3 1
A proof of Eq. [5] is given in Refs. (4, 29). Equation exp兵c其 ⫽ 1 ⫹ c ⫹ c ⫹ c ⫹ · · · ⫹ cN ⫹ · · ·
2! 3! N!
[6] may be expressed in the Dirac formalism as a sum
of products of eigenvalues and projection operators. [12]

冘 a 兩j典具j兩
ᏺ an exponential operator may formally be interpreted
 diag ⫽ j [7] as
j⫽1

1 1
where each operator 兩j典具j兩 projects a given state vector exp兵iÂ其 ⫽ 1̂ ⫹ Â ⫹ 共iÂ兲2 ⫹ 共iÂ兲3
2! 3!
onto the eigenstate 兩j典 (2– 4, 6, 29). We may check
that Eqs. [6] and [7] are equivalent by calculating the 1
⫹···⫹ 共iÂ兲N ⫹ · · · [13]
kth element of Âdiag, that is, (Âdiag)kk ⫽ 具k兩Âdiag兩k典. N!
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 121

where Â2 ⫽ Â 䡠 Â, and so forth. How is a power of an exp{iÂdiag} is represented by a diagonal matrix with
operator ÂN calculated for large N? Equation [5] is exponentials of eigenvalues on the diagonal:
helpful because, for example, it may be exploited to
obtain Â3 as follows: exp兵iÂdiag其

冢 冣
exp兵ia1 其 0 0 ··· 0
 3 ⫽  䡠  䡠  ⫽ 共X̂ÂdiagX̂† 兲共X̂ÂdiagX̂† 兲共X̂ÂdiagX̂† 兲
0 exp兵ia2 其 0 ··· 0
[14] ⫽ 0 0 exp兵ia3 其 ··· 0
·· ·· ·· ·· ··
· · · · ·
The right-hand side may be simplified by exploiting 0 0 0 · · · exp兵iaᏺ 其
that X̂ is unitary and hence that the products X̂†X̂
[20]
cancel (Eq. [4]). Equation [14] then casts as
Inserting this result into Eq. [19] gives the following
 3 ⫽ X̂Âdiag
3
X̂† [15] expression for the exponential operator exp{iÂ}:

Note that whenever Âdiag is diagonal, it may be di- exp兵iÂ其 ⫽ X̂ exp兵iÂdiag其X̂† [21]
rectly raised to any power N according to
If the operator  is Hermitian, the corresponding

冢 冣
a1N 0 0 ··· 0 exponential operator exp{iÂ} is unitary (6, 29). It
0 a2N 0 ··· 0 follows that the inverse of the exponential operator is
N
N
 diag ⫽ 0 0 a3 ··· 0 [16] given by exp{iÂ}⫺1 ⫽ exp{⫺iÂ}. This may be ver-
·· 0 0 ·· ·
· · ·· ified by exploiting Eqs. [1] and [4] as follows:
0 0 0 · · · aᏺN
exp兵iÂ其⫺1 ⫽ 共exp兵iÂ其兲† ⫽ exp兵共i兲† † 其 ⫽ exp兵⫺iÂ其
From this follows that Â3 is formed by first calculat-
ing Â3diag and then multiplying it with X̂ from the left [22]
and X̂† from the right.
When applying this strategy to Eq. [13], we obtain Equation [21] is the route for obtaining exponential
operators in numerical applications. The procedure to
1 calculate exp{iÂ} is as follows:
exp兵iÂ其 ⫽ 1̂ ⫹ 共iX̂ÂdiagX̂† 兲 ⫹ 共iX̂ÂdiagX̂† 兲2 ⫹ · · ·
2!
1. diagonalize operator  (i.e., find its eigenvalues
1 and eigenvectors),
⫹ 共iX̂ÂdiagX̂† 兲N ⫹ · · · [17]
N! 2. construct the matrix X̂ having the eigenvectors
of  as columns,
1 3. calculate the matrix of complex exponentials of
⫽ 1̂ ⫹ 共iX̂ÂdiagX̂† 兲 ⫹ 共iX̂Âdiag
2
X̂† 兲 ⫹ · · · the eigenvalues (Eq. [20]), and
2!
4. form the product X̂ exp{iÂdiag}X̂† (Eq. [21]).
1
⫹ 共iX̂Âdiag
N
X̂† 兲 ⫹ · · · [18]
N!
Fourier Series and Fourier Transforms
Because each term is “sandwiched” by X̂ from the left This section discusses basic concepts that are impor-
and by X̂† from the right, the sum may be factorized tant when processing NMR responses, as well as in
as the mathematical formulation of the time evolution of
nuclear spin systems. Here we also define parameters


exp兵iÂ其 ⫽ X̂ 1̂ ⫹ iÂdiag ⫹
1
2!
共iÂdiag兲2
relevant for the experimental acquisition of NMR
time-domain signals.

⫹· · · ⫹
1
N!
共iÂdiag兲N ⫹ · · · X̂† 冊 [19]
Fourier Series. A real- or complex-valued function
f(t) is said to be periodic with a characteristic period
T, if it obeys f(t) ⫽ f(t ⫹ T) for all values of t. Here
The sum in the parentheses may be identified, accord- we assume that the variable t represents time, which is
ing to Eq. [13], as the Taylor expansion of relevant for spin dynamics calculations and in the
exp{iÂdiag}. In analogy with Eq. [16], the operator acquisition of an experimental NMR time-domain
122 EDÉN

signal. The period T is associated with a modulation of g(t) generally comprise an infinite number of Fou-
frequency ␻mod, which is inversely proportional to T: rier components g (k) .

␻ mod ⫽ 2␲/T [23] Fourier Transforms. The Fourier transform (FT) is


used extensively in the processing of NMR experi-
Since T is given in units of seconds, ␻mod is an ments: it converts an NMR time-domain signal s(t)
angular frequency given in rads per second. A fre- (amplitude as a function of time) into a frequency-
quency ␯ (Hz) is related to the corresponding angular domain spectrum S(␻) (amplitude as a function of
frequency ␻ by frequency). The actual conversion is defined mathe-
matically as an integration of the product
␻ ⫽ 2␲␯ [24] s(t)exp{⫺i␻t} over all time points t

A periodic function may be expanded in a Fourier ⬁

series
S共␻兲 ⫽ 共2␲兲 ⫺1 冕 dt S共t兲exp兵⫺i␻t其 [28]


M/ 2

f共t兲 ⫽ f 共k兲exp兵ik␻modt其 [25] ⫺⬁

k⫽⫺M/ 2
A physical interpretation of the Fourier transform
where the time-independent coefficients f (k) are called procedure may be found in Ref. (6). This operation is
Fourier components (or Fourier coefficients). Equa- also invertible; the inverse Fourier transform (IFT) is
tion [25] shows that f(t) may be expressed as a sum defined as
over M ⫹ 1 Fourier components, each weighted by a
complex exponential function involving a harmonic ⬁
of ␻mod. The number of Fourier components required
to reproduce the function f depends on the nature of
its time dependence. In general, the more complicated
s共t兲 ⫽ 冕 d␻ S共␻兲exp兵i␻t其 [29]

dependence on time, the larger the number M. The ⫺⬁

value of M may sometimes extend to infinity. The


Fourier components of a periodic function f may be It converts the frequency-domain spectrum S(␻) into
calculated from the relationship the time-domain signal s(t) by an integration of the
product S(␻)exp{i␻t} over all frequencies ␻. The
T interconversion between s(t) and S(␻) by the Fourier


transform and inverse Fourier transform operations is
1
f 共k兲 ⫽ dt f共t兲exp兵⫺ik␻modt其; illustrated in Fig. 2 for the case of a simple NMR
T signal.
0
Discrete Fourier Series and Fourier Transforms. In
M M M practice, the experimentally acquired NMR time-do-
k ⫽ ⫺ , ⫺ ⫹ 1, . . . , [26]
2 2 2 main signal is not continuous, but corresponds to a set
of q discrete time points t j and amplitudes s(t j ): {t j ,
In the theoretical formalism outlined in later sec- s(t j )}. All information known about the function s(t)
tions, we will encounter complex exponentials of pe- is contained in its q discrete values. A discretely
riodic functions, such as g(t) ⫽ exp{if(t)}. Such sampled time signal is illustrated in Fig. 3. For prac-
exponential functions are also periodic, that is, tical reasons q is assumed to be even. The signal is
exp{if(t ⫹ T)} ⫽ exp{if(t)}, and may consequently sampled over an acquisition interval ␶acq at the
be expanded in a Fourier series, similar to Eq. [25]: equally spaced time points

t j ⫽ j␶ acq/q, j ⫽ 0, 1, 2, . . . , q ⫺ 1
冘g
⬁ [30]
共k兲
exp兵if 共t兲其 ⫽ exp兵ik␻modt其 [27]
k⫽⫺⬁ The duration between two time points t j and t j⫹1 is
usually called the dwell time (note, however, that it is
Note, however, that even in the case of a finite number a time interval) and may be calculated from either of
of Fourier components in Eq. [25], the Fourier series the two following relationships:
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 123

␻ res ⫽ ␻samp /q [34]

or, alternatively, from the inverse of the acquisition


interval

␻ res ⫽ 2␲/␶acq [35]

Note that ␻res has the same role in the frequency


domain as the “dwell time” (␶dwell) has in the time
domain. The two entities are calculated from ␶acq and
␻samp using closely related expressions: Eq. [31] is

Figure 2 (a) The NMR time-domain signal is converted


into (b) the frequency-domain spectrum by a Fourier trans-
form (FT). The reverse transformation is effected by an
inverse Fourier transform (IFT). In this case, the time-
domain signal is a sum of two components, shown in (c) and
(e), respectively. The corresponding spectra of the signals
are shown in (d) and (f), respectively. Note that (c) the
component with the higher oscillation frequency corre-
sponds to (d) the peak at the higher frequency in the spec-
trum. Generally, the NMR signals are complex, but for
simplicity, we have only displayed their real parts.

␶ dwell ⫽ ␶acq/q [31]

⫽ 2␲/␻ samp [32]

where the sampling frequency ␻samp corresponds to


the span of spectral frequencies (see Fig. 3).
The spectrum is obtained by a discrete Fourier
transformation procedure (discussed below) and com-
prises the same number of points as does the time-
domain signal. It corresponds to a set of frequency-
Figure 3 The relationship between parameters used when
amplitude pairs {␻ k , a k }. The expression for the acquiring an NMR time-domain signal. (a) The signal is
amplitudes a k are defined below, and the spectral acquired at q discrete time points (in the present case, q ⫽
coordinates are displaced around ␻ ⫽ 0 and range 32) with a duration ␶dwell between them. This results in a
from the negative frequency ⫺␻samp /2 ⫹ ␻res to the total time span ␶acq, called the acquisition interval. (b)
positive frequency ⫺␻samp /2. The qth frequency co- Subsequent discrete Fourier transformation results in the
ordinate is (rad s⫺1) given by NMR frequency-domain spectrum, consisting of q frequency–
amplitude pairs. This extends over the frequency range
q q q ␻samp, which is inversely proportional to ␶dwell (Eq. [32]).
␻ k ⫽ k␻ res; k ⫽ ⫺ ⫹ 1, ⫺ ⫹ 2, . . . , [33] The frequencies are spaced evenly in steps of the frequency
2 2 2
resolution ␻res, ranging from the most negative frequency
(left side of spectrum) ␻ ⫽ ⫺␻samp /2 ⫹ ␻res to the most
The frequency-domain resolution ␻res corresponds to positive value ␻ ⫽ ␻samp /2 (right side of spectrum). The
the separation between two neighboring spectral co- frequency resolution is inversely proportional to the signal
ordinates: ␻res ⫽ ␻ k⫹1 ⫺ ␻ k . It may be calculated by acquisition interval ␶acq (Eq. [35]). The lines between the
dividing the spectral range ␻samp by the number of points are only for visualization purposes and have no
points q physical significance.
124 EDÉN

analogous to Eq. [34], whereas Eq. [32] is analogous operator for a rotation by an angle ␪ around an axis nᠬ
to Eq. [35]. is denoted R̂n (␪ ) and defined as (28, 29)
From Eq. [35] it follows that the longer the exper-
imental acquisition interval, the finer the spectral res- ᠬ
R̂ n 共␪兲 ⫽ exp兵⫺i␪l̂ 䡠 nᠬ 其 [38]
olution. Note that the convention used here for the
spectral range ␻samp includes the positive frequency
␻ ⫽ ⫹␻samp /2 but not the corresponding negative where ᠬl̂ is the total angular momentum operator.
value ␻ ⫽ ⫺␻samp /2. In fact, the amplitudes at the A rotation of any 3-dimensional object can be
extreme positive and negative coordinates (i.e., a ⫺q/ 2 specified by a sequence of three consecutive single-
and a q/ 2 ) are, by definition, equal. This follows from axis rotations and may, therefore, be parameterized by
the properties of the Fourier transform of discrete data three axes and three corresponding angles. Here we
sets and is discussed in detail in Ref. (30). However, make use of passive rotations, which are transforma-
in order to keep the same number of points (q) in both tions of the coordinate systems rather than the objects
the time and frequency domains, our convention is themselves (4, 28, 29). The transformation of the
simply to drop the most negative frequency coordi- coordinate system F with basis vectors {xᠬ F , yᠬ F , zᠬF }
nate and use a slightly asymmetrical spectral range. into the system F⬘ ⫽ {xᠬ F⬘ , yᠬ F⬘ , zᠬF⬘ } may be effected
This has negligible consequences in practice, as long by the following sequence of rotations:
as the frequency resolution ␻res is fine enough.
We point out that s(t) is generally not periodic in R̂共␣FF⬘ , ␤FF⬘ , ␥FF⬘ 兲 ⫽ R̂z F⬘共␥FF⬘ 兲R̂y G共␤FF⬘ 兲R̂z F共␣FF⬘ 兲
time. However, functions that are sampled as discrete [39]
points t j may always be expressed as a Fourier series
according to Eq. [25]. References (4) and (30) explain Equation [39] may be interpreted as follows: first the
the reasons for this property of discrete functions in coordinate system is rotated around its zᠬF axis by an
more detail. angle ␣ FF⬘ . Then it is rotated around the y axis (here
The corresponding set of Fourier components {a k } denoted yᠬ G ) of the new system by an angle ␤ FF⬘ ,
are obtained by a discrete Fourier transformation (30) followed by a rotation around the z axis of the “final”
of the set {s(t j )} as follows: coordinate system F⬘ by an angle ␥ FF⬘ . The rotation
is therefore parameterized by a triplet of Euler angles

冘 s共t 兲exp兵⫺i2␲jk/q其;
q⫺1
[using the convention given on p. 21–22 of Ref. (28)]
a k ⫽ q ⫺1 j
j⫽0
⍀ FF⬘ ⫽ 兵␣ FF⬘, ␤ FF⬘, ␥ FF⬘其 [40]
q q q
k ⫽ ⫺ ⫹ 1, ⫺ ⫹ 2, . . . , [36] where the angles take the values {0 ⱕ ␣ FF⬘ ⬍ 2␲,
2 2 2
0 ⱕ ␤ FF⬘ ⱕ ␲, 0 ⱕ ␥ FF⬘ ⬍ 2␲}. Note that the
The Fourier component a k corresponds to the spectral subscripts FF⬘ specify a transformation from system
amplitude at the frequency coordinate ␻ k . Note that in F to system F⬘ (although intermediate systems are
Eq. [36] the index j runs over all integers 0 ⱕ j ⱕ involved in each single-axis rotation). However, it
q ⫺ 1 whereas the index k takes integer values in the may be shown (28) that the same transformation can
range ⫺(q/ 2) ⫹ 1 ⱕ k ⱕ q/ 2. be carried out by employing rotations solely around
As in the case of continuous functions, a discrete the axes of the original coordinate system F, if the
inverse Fourier transform is defined to produce s(t j ) order of rotations is reversed:
from the set of spectral amplitudes {a k } as follows:
R̂共␣FF⬘ , ␤FF⬘ , ␥FF⬘ 兲 ⫽ R̂z F共␣FF⬘ 兲R̂y F共␤FF⬘ 兲R̂z F共␥FF⬘ 兲

q/ 2

s共t j兲 ⫽ a kexp兵i2␲jk/q其 [37] [41]


k⫽⫺q/ 2⫹1

We stress that we used angular frequencies for the Only two Euler angles are needed for specifying
spectral coordinates in this section. If desired, these transformations of objects being symmetric around
may be converted into Hertz using Eq. [24]. one axis, such as vectors. Equation [41] then reduces
to a rotation by the angle ␤ FF⬘ around the y F axis,
followed by a rotation by the angle ␣ FF⬘ around the z F
Euler Angles and Rotation Operators
axis. This is illustrated in Fig. 4. In this case, the
A counterclockwise rotation around an axis is defined following relationship holds between the Euler angles
as positive and a clockwise rotation as negative. The {␣, ␤} and the polar angles {␾, ␪}:
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 125

␣⬅␾ [42]

␤⬅␪ [43]

As discussed in a following publication, this relation-


ship is useful for visualizing the orientational depen-
dence of NMR interactions, especially when calculat-
ing powder averages.
In quantum mechanics, Eq. [41] is presented in
terms of the Wigner rotation operator (28, 29).

D̂共␣FF⬘ , ␤FF⬘ , ␥FF⬘ 兲


⫽ exp兵⫺i␣FF⬘ l̂z 其exp兵⫺i␤FF⬘ l̂y 其exp兵⫺i␥FF⬘ l̂z 其 [44]

where we used the shorthand notation x ⫽ x F , y ⫽


y F , and z ⫽ z F . The rotation operators are usually
represented in a basis composed of a set of 2l ⫹ 1
simultaneous eigenstates {兩lm典} of the operator for
the square of the total angular momentum ᠬl̂ 2 and the
operator l̂z (28, 29):

l̂ᠬ 2 兩lm典 ⫽ l共l ⫹ 1兲兩lm典 [45]

l̂ z 兩lm典 ⫽ m兩lm典; m ⫽ ⫺l, ⫺l ⫹ 1, . . . , l [46]

The “eigenequations” involve the quantum number l


for the total angular momentum and its component m
along the z direction (28, 29). The matrix elements
l
D m⬘m of the lth rank rotation operator are then given
by

l
D m⬘m共⍀ FF⬘兲 ⬅ 具lm⬘兩D̂共⍀FF⬘ 兲兩lm典

⫽ 具lm⬘兩exp兵⫺i␣FF⬘ l̂z 其exp兵⫺i␤FF⬘ l̂y 其

⫻ exp兵⫺i␥FF⬘ l̂z 其兩lm典 [47]


Figure 4 The orientation of an object that has a symmetry
axis (e.g., a vector) may be parameterized by the two Euler
where m and m⬘ may take all integer values between
angles {␣, ␤}, or equivalently, by the two polar angles {␪,
⫺l and l, giving a total of (2l ⫹ 1) ⫻ (2l ⫹ 1) such
␾}. Assume that the vector is aligned along the z F axis of
the coordinate system F (whose axes are labeled in the Wigner elements. It is convenient to define a reduced
figure). According to Eq. [41], the vector may be rotated to Wigner element (28, 29) by
the orientation {␣, ␤} [indicated by the circle on the surface
of the sphere shown in (c)] by a sequence of three rotations: l
d m⬘m共␤ FF⬘兲 ⬅ 具lm⬘兩exp兵⫺i␤FF⬘ l̂y 其兩lm典 [48]
first, by the angle ␥ around the z F axis; second, by the angle
␤ around the y F axis; and third, by the angle ␣ around the l
and express D m⬘m (⍀ FF⬘ ) as
z F axis. However, as a rotation around a symmetry axis
leaves the vector unaffected, the first transformation around
the z F axis does not cause any net rotation. The second and
l
D m⬘m共⍀ FF⬘兲 ⫽ exp兵⫺im⬘␣FF⬘ 其dm⬘m
l
共␤FF⬘ 兲exp兵⫺im␥FF⬘ 其
third rotations give an orientation that in this case corre- [49]
sponds to the orientation {␣, ␤} ⫽ {␲/2, ␲/4}, or in terms
of polar angles, {␪, ␾} ⫽ {␲/4, ␲/2}. The reduced Wigner elements are real-valued func-
tions of ␤. Explicit expressions for d m⬘m
1 2
and d m⬘m are
given in Table 1.
126 EDÉN

Table 1 Reduced Wigner Functions


1
d m⬘m (␤)
m⬘\ m 1 0 ⫺1

1 1 1
1 2
(1 ⫹ cos ␤) ⫺ 冑 2 sin ␤ 2
(1 ⫺ cos ␤)
1 1
0 冑 2 sin ␤ cos ␤ ⫺ 冑2 sin ␤
1 1 1
⫺1 2
(1 ⫺ cos ␤) 冑 2 sin ␤ 2
(1 ⫹ cos ␤)

2
d m⬘m (␤)

m⬘\ m 2 1 0 ⫺1 ⫺2

2 1
4
(1 ⫹ cos ␤)2 ⫺21 sin ␤ (1 ⫹ cos ␤) 冑 3
8
sin2␤ ⫺1 sin ␤ (1 ⫺ cos ␤)
2
1
4
(1 ⫺ cos ␤)2

1 1
2
sin ␤ (1 ⫹ cos ␤) 1
2
(2 cos2␤ ⫹ cos ␤ ⫺ 1) ⫺ 冑 3
8
sin 2␤ 1
2
(2 cos2␤ ⫺ cos ␤ ⫺ 1) ⫺1 sin ␤ (1 ⫺ cos ␤)
2

0 冑 3
8
sin2␤ 冑 3
8
sin 2␤ 1
2
(3 cos2␤ ⫺ 1) ⫺ 冑 3
8
sin 2␤ 冑 3
8
sin2␤

⫺1 1
2
sin ␤ (1 ⫺ cos ␤) ⫺21 (2 cos2␤ ⫺ cos ␤ ⫺ 1) 冑 3
8
sin 2␤ 1
2
(2 cos2␤ ⫹ cos ␤ ⫺ 1) ⫺1 sin ␤ (1 ⫹ cos ␤)
2

⫺2
1
4
(1 ⫺ cos ␤)2 1
2
sin ␤ (1 ⫺ cos ␤) 冑 3
8
sin2␤ 1
2
sin ␤ (1 ⫹ cos ␤) 1
4
(1 ⫹ cos ␤)2

Irreducible Spherical Tensors irreducible spherical tensor. Consequently, three basis


“functions” are necessary to define the vector: in the
Many objects have properties that depend on their
present case, those are the three Cartesian orthogonal
orientation with respect to measurements or certain
unit vectors, and each vector component v j is given by
manipulations. For example, a piece of wood or meat
the projection of vᠬ onto the corresponding unit vector.
has fibers in certain directions, which makes it more
For example, the component v z is calculated as the
difficult to cut across the direction of the fibers than
scalar product of vᠬ and zᠬ : v z ⫽ vᠬ 䡠 zᠬ.
parallel to them. This type of “anisotropic behavior”
The irreducible spherical tensor is defined by its
is closely related to the symmetry of the object with
response to rotations of the coordinate system, which
respect to the operation performed upon it.
“Anisotropic” objects are conveniently described
in terms of irreducible spherical tensors. An irreduc-
ible spherical tensor of rank l, expressed in frame F,
is composed of 2l ⫹ 1 elements,

7
关A 共l 兲 兴F ⫽ 共关All 兴F , 关All⫺1 兴F , . . . , 关Al⫺l 兴F 兲 [50]

with the mth component denoted [A lm ] F and in gen-


eral represented by a complex number [Fig. 5(a)].
Note that the number of components of the irreducible
spherical tensor depends on its rank. Components
with opposite signs of the index m are related through

关A lm兴 F ⫽ 共⫺1兲m 关Al⫺m 兴F * [51]


Figure 5 (a) A second rank (l ⫽ 2) irreducible spherical
tensor has five components A lm . Each of these is in a
where the asterisk denotes complex conjugation. Each
general reference frame represented by a complex number,
component corresponds to the projection of the tensor except for A l0 which is real. The components A lm and A l⫺m
onto a certain basis function (4, 28, 29); to define an are related through Eq. [51]. (b) The components of a
lth rank tensor, 2l ⫹ 1 basis functions are required. second rank irreducible spherical tensor operator. The com-
For example, consider a 3-dimensional vector repre- ponents T̂lm correspond to spin operators, represented as
sented in a Cartesian coordinate system vᠬ ⫽ (v x , v y , matrices. This example assumed operators for two coupled
v z ). Such a vector corresponds to a first-rank (l ⫽ 1) spins-21.
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 127

is completely specified by its rank l. Two irreducible Sometimes the components of the spherical tensors
spherical tensors of the same rank respond similarly to are not complex numbers but operators [Fig. 5(b)].
rotations, but these responses are different from that An irreducible
7
spherical tensor operator of rank l is
of another irreducible spherical tensor of different denoted T̂(l ) . Each of its 2l ⫹ 1 components may be
rank. An lth rank irreducible spherical tensor is trans- represented as a matrix in some suitable basis set, and
formed from frame F to frame F⬘ by multiplication it has the following symmetry upon sign reversal of m
with the Wigner rotation operator of the same rank, (28, 29):
according to
T̂ lm ⫽ 共⫺1兲m T̂l⫺m

[56]
7 7 共l 兲
关A 兴 ⫽ 关A 兴 D̂ 共⍀FF⬘ 兲
共l 兲 F⬘ 共l 兲 F [52]

It follows that the mth component of the tensor in the NMR INTERACTIONS IN SOLID STATE
“new” frame is related to the components of the “old”
frame according to Overview
The Zeeman interaction, which is the coupling of the

l
spin angular momentum to the static magnetic field
关A lm兴 F⬘ ⫽ 关A lm⬘兴 FD m⬘m
l
共⍀ FF⬘兲 [53]
m⬘⫽⫺l
ជ , is characterized by the Larmor frequency
B0

Hence, under rotations, the values of the individual ␻ 0 ⫽ ⫺␥I B0 [57]


components change but the number of components is
fixed. For example, the vector vᠬ is always completely where ␥ I is the magnetogyric ratio (1– 6) of the spins
specified by exactly three components. This is the and B 0 is the magnitude of the field. However, each
same as stating that the rank of an irreducible spher- spin in the sample senses a slightly different magnetic
ical tensor is conserved upon rotation, in the same field (“local field”) at its location, because of its very
way as the shape of a rigid object remains unchanged specific chemical environment. The spin interactions
when the object is rotated. An object that responds arising from these local fields may be classified into
equally in all directions is said to be isotropic. It is, chemical shift and spin–spin interactions (1– 6). The
therefore, invariant to rotations and may be repre- spin–spin interactions that reflect couplings between
sented as a zeroth rank (l ⫽ 0) tensor (a “scalar”). As the spins may be categorized further into homo-
discussed below, Eqs. [52] and [53] are used exten- nuclear couplings (involving spins of the same spe-
sively for transforming NMR interactions between cies, e.g., two 1H spins) and heteronuclear couplings
various coordinate systems. (involving spins of different species, e.g., a 1H and a
Equation [52] may be viewed as a vector–matrix 13
C). Also, depending on the mechanism of the cou-
multiplication: the irreducible spherical tensor repre- plings, the spin–spin interactions may be divided into
sented in frame F⬘ is obtained by multiplying a row through-space dipole– dipole (henceforth referred to
vector of dimension 2l ⫹ 1 with a matrix of dimen- as dipolar interactions) and through-bond (scalar) in-
sion (2l ⫹ 1) ⫻ (2l ⫹ 1). This interpretation is teractions (henceforth referred to as J interactions).
conveniently exploited in computer programs, espe- Quadrupolar nuclei, having spin number I ⬎ 1/ 2, are
cially if several consecutive transformations are also affected by first- and second-order quadrupolar
needed. For example, the transformations F 3 F⬘ 3 interactions (33).
F⬙ may be carried out through

7 7
关A 共l 兲 兴F⬙ ⫽ 关A共l 兲 兴F D̂共l 兲 共⍀FF⬘ 兲D̂共l 兲 共⍀F⬘F⬙ 兲 [54]
NMR Interactions in Terms of Irreducible
Spherical Tensors
Note carefully the ordering of the subscripts. An important feature of NMR interactions is their
The scalar product (inner product) of two irreduc- dependence on the orientation of the molecule rela-
ible spherical tensors is used extensively when con- tive to the direction of the static magnetic field, im-
structing Hamiltonians in NMR. It is defined by plying that the NMR spectrum of a specific nuclear
site in a molecule depends on the molecular orienta-

冘 共⫺1兲 关A 兴 关A
l tion. This will be discussed in detail in a following
7 7
关A J共l 兲 兴F 䡠 关A共lK 兲 兴F ⫽ m J F
lm
K
l⫺m兴F
[55] publication. The interactions may be separated into
m⫽⫺l parts that are anisotropic (orientation dependent) and
128 EDÉN

Table 2 Construction of Spin Hamiltonians from Spatial and Spin Components of Various Interactions
Interaction C⌳ A⌳
00 T̂⌳
00
iso
Ĥ⌳ [A ⌳
20 ]
L
[T̂⌳
20 ]
L aniso
Ĥ⌳

Chemical shift (CS) ⫺␥ I ⫺ 公3 ␦iso


j

1
冑3 B0Îjz ␦iso
j
␻ 0 Îjz ␦CSA
j
冑 2
3
B 0 Îjz ␻CSA
j
冑 2
3
Îjz
Homonuclear 1 1
␻D
冑6 (3ÎjzÎkz ⫺ Îj 䡠 Îk) ␻ D 冑6 (3ÎjzÎkz ⫺ Îj 䡠 Îk)
jk jk
dipolar (D) 1 — — —

冑 冑
Heteronuclear
2 2
dipolar (D) 1 — — — ␻D
IS
Îz Ŝz ␻D
IS
Îz Ŝz
3 3
1 1 1
⫺2␲ 公3 J iso ⫺
冑3 Îj 䡠 Îk 2␲J isoÎj 䡠 Îk ␻ J,aniso 冑6 (3ÎjzÎkz ⫺ Îj 䡠 Îk) ␻ J,aniso 冑6 (3ÎjzÎkz ⫺ Îj 䡠 Îk)
jk jk jk jk
Homonuclear (J) 1

Heteronuclear (J) 1 ⫺2␲ 公3 J iso


IS

1
冑3 ÎzŜz
IS
2␲J iso Îz Ŝz ␻ J,aniso
IS
冑 2
3
Îz Ŝz ␻ J,aniso
IS
冑 2
3
Îz Ŝz
First-order 1 1
1
[Q j20 ] L
冑6 (3Îjz ⫺ I(I ⫹ 1)) ␻ Q 冑6 (3Îjz ⫺ I(I ⫹ 1))
2 j 2
quadrupolar (Q) — — —
2I(2I ⫺ 1)

The spin interaction Hamiltonians may be constructed from Eq. [20] as the product of a constant C ⌳ , the spatial part in the laboratory frame
⌳ L ⌳ L
[A l0 ] , and the spin part [T̂l0 ] . (See the discussion in the text regarding the chemical shift Hamiltonian.) The contribution from the l ⫽ 0
part of interaction ⌳ is obtained by multiplying columns 2, 3, and 4 of the given row. The result is given in column 5 (Ĥ⌳ iso
). The contribution
from the l ⫽ 2 part is obtained by multiplication of columns 2, 6, and 7, resulting in the expression given in column 8 (Ĥ⌳ aniso
). The tensor
⌳ P
components [A 2m ] necessary for obtaining [A ⌳ L
20 ] are listed in Table 3. The various parameters and constants have the following meaning:
for spin species I, ␥ I denotes the gyromagnetic ratio and I is the spin number (i.e., total spin angular momentum). The isotropic components
A iso (e.g. ␦iso
J IS
and J iso ) are defined as the mean value of the principal values of the spatial tensor:

1
A iso ⫽ 共关Axx 兴P ⫹ 关Ayy 兴P ⫹ 关Azz 兴P 兲.
3

For the first-order quadrupolar interaction, the quadrupolar frequency is defined as

关Q j20兴 L
␻ Qj ⫽
2I共2I ⫺ 1兲

isotropic (orientation independent). As explained fur- parts couple together to form what we refer to as the
ther below, the isotropic parts correspond to the av- NMR “interaction.”
erage of the interaction over all orientations of the
molecule. However, in the cases of through-space
dipolar and first-order quadrupolar interactions, the General Form of Spin Hamiltonian
isotropic parts are zero.
The spin Hamiltonian representing the interaction ⌳
It is convenient to describe the responses of the
may be expressed as a sum of scalar products of
NMR interactions to molecular rotations by means of
irreducible spherical tensors, where each scalar prod-
irreducible spherical tensors: each NMR interaction
uct involves a spatial tensor and a spin tensor (1–5).
may be represented by a zeroth rank (l ⫽ 0) or
The irreducible spherical tensors may be expressed in
second rank (l ⫽ 2) irreducible spherical tensor. The
different frames of reference. Because the Hamilto-
second-order quadrupolar interaction is a special case.
nian is a scalar product, it is independent of the choice
We will not consider it here, but refer to Ref. (33) for
of frame of the tensors (as long as they are represented
details about this interaction.
in the same frame). However, the laboratory frame
NMR interactions may be thought of as deriving
(L) is the natural choice. It is defined such that its z
from two parts: one is dependent on “molecular prop-
axis coincides with the direction of the static magnetic
erties” and is henceforth referred to as the “spatial
field. The laboratory frame Hamiltonian of interaction
tensor.” The lth rank part of this tensor is denoted
7(l ) ⌳ may be expressed (1–5) as
A⌳ , and it may be represented as a set of 2l ⫹ 1

complex numbers A lm as depicted in Fig. 5(a). The
other, referred to as the “spin part,”
7
involves an irre-
(l )
ducible spherical tensor operator T̂ ⌳ , which depends Ĥ ⌳ ⫽ C⌳ 冘 关A 7 共l 兲 L

7
兴 䡠 关T̂ ⌳共l 兲 兴L [58]
on spin angular momentum. The elements T ᠬ lm

corre- l⫽0,2

spond to spin operators that themselves are repre-


sented by matrices [Fig. 5(b)]. Together these two Using Eq. [55], this may be written
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 129

izations and additionally first rank with respect to


冘 冘
l
⌳ L ⌳
Ĥ ⌳ ⫽ C⌳ 共⫺1兲m 关Alm 兴 䡠 关T̂l⫺m 兴L [59] rotations of the external magnetic field. The spin-field
l⫽0,2 m⫽⫺l tensor may be of both zeroth and second rank with
respect to simultaneously rotating the spins and the
where C ⌳ is a constant characteristic of the interac- direction of the external magnetic field. This is the
tion. reason for the appearance of the factors ⫺(1/ 公3)
In high-field NMR (ignoring second-order quadrupo- B 0 Îz and 公2/3 B 0 Îz in Table 2. Because the magnetic
lar interactions in the case of quadrupolar nuclei), one field direction is never changed in conventional NMR
may show that the only significant parts of the Hamil- experiments, it is often practical to combine C CS, B 0 ,
7
tonian are those whose spin parts commute with the the and the relevant component of ␦CS together. Loosely,
operator for the z component of the total spin angular this will be referred to as a “chemical shift tensor in
7(l )
momentum Îz. The commuting tensor operator compo- frequency units,” denoted ␻CS . Starting from its def-
nents are T̂⌳l0 (1–5). Equation [59] then simplifies to
inition as a product of a chemical shift tensor com-
ponent and a spin-field component, the Hamiltonian of
⌳ ⌳ ⌳ L ⌳ L the isotropic chemical shift interaction of spin j may be
Ĥ ⌳ ⫽ C⌳ 兵A00 T̂00 ⫹ 关A20 兴 关T̂20 兴其 [60]
rearranged as
Note that because the zeroth rank tensors are invariant
to rotations, it is not necessary to specify the partic-
j
Ĥ CS,iso ⫽ ⫺␥I ␦ 00
j
Ᏺ̂ 00
j
[61]
ular choice of reference frame for these. Equation [60]
furnishes a framework for constructing Hamiltonians
of spin interactions. The caption of Table 2 demon-
strates how this may be done in practice for the
⫽ ⫺␥I 共⫺ 冑3␦iso
j
兲 ⫺ 冉 1
冑3
B0 Îjz 冊 [62]

chemical shift, dipolar, J, and first-order quadrupolar


Hamiltonians. All Hamiltonians are given in angular ⫽ ␦ iso
j
␻0 Îjz [63]
frequency units (rad s⫺1), as can be verified from the
interaction parameters listed in Table 2. using the expressions from Table 2 and the definition
We emphasize that the high-field spin Hamiltonian is of the Larmor frequency (Eq. [57]). Equation [63] is
a product of two components of irreducible spherical the “conventional” form of the chemical shift Hamil-
tensors, which have to be of the same rank l. In most tonian usually found in the literature (1– 6).
cases they may be directly interpreted as a “spatial part” In summary, care has to be taken when forming
(which reflects how the interaction changes when the spin Hamiltonians from irreducible spherical tensors.
molecule is rotated) and a spin part (which reflects how However, because the Hamiltonian is crucial when
the interaction changes upon rotations of the spin polar- writing numerical simulation programs, it is beneficial
izations). Take, for example, the anisotropic hetero- to exploit the irreducible spherical tensor concept,
IS
nuclear J-coupling Hamiltonian ĤJ,aniso . It involves a because it provides a general framework for system-
component of the J tensor, which is second rank with atic construction of Hamiltonians, as well as for the-
respect to rotations of the molecule. The spin part (公2/3 oretical analyses of NMR experiments.
ÎzŜz) is second rank with respect to a simultaneous
rotation of both the I and S spin polarizations. However,
it is first rank with respect to rotations of either the I or
S spin polarizations alone. SPIN OPERATORS
The classification of spatial and spin parts of the
Hamiltonian is usually straightforward but may in some Zeeman Basis
cases be obscure. This is especially the case for the
The tensor operator components T̂⌳ l0 appearing in the
chemical shift interaction involving the coupling of spins
expression of the Hamiltonian (Eq. [60]) are linear
to the external magnetic field 共B ជ兲 through the chemical
7
0 combinations of spin operators. In a numerical simu-
shift (CS) tensor ␦CS, which may be decomposed lation program, the explicit matrix representations are
into parts with l ⫽ 0, 1, or 2. The relevant ones to needed for these operators. The relationship between
NMR are either zeroth or second rank with respect to the T̂⌳l0 and the spin product operators for various
rotations of the molecule. Strictly speaking, the “spin NMR interactions are given in Table 2. The operators
parts” of the chemical shift interaction given in Table are normally represented in the Zeeman basis, which
7
2 are components of the spin-field tensor Ᏺ̂. This is is the eigenbasis of Îz . Here we outline how this basis
first rank with respect to rotations of the spin polar- may be constructed for a system of coupled spins-21.
130 EDÉN

The basis set for systems of quadrupolar nuclei (I ⬎ This process may be used recursively for con-
1/ 2) is constructed analogously. structing the Zeeman basis states for larger spin sys-
A single spin-21 has two states of well-defined spin tems. For example, the three-spin product state 兩␤␤␣典
angular momentum along the static field direction. is, according to Eq. [66], calculated as 兩␤␤␣典 ⫽ 兩␤典 R
They are called Zeeman states: 兩␣典 denotes the state 兩␤典 R 兩␣典. The direct product operation is associative,
with the quantum number for total spin angular mo- meaning that we may form the state 兩␤␤␣典 by two
mentum I ⫽ 1/ 2 and z component m ⫽ 1/ 2, whereas consecutive products, as either 兩␤␤␣典 ⫽ 兩␤典 R (兩␤典 R 兩␣典)
兩␤典 is the state with I ⫽ 1/ 2 and z component m ⫽ or 兩␤␤␣典 ⫽ (兩␤典 R 兩␤典) R 兩␣典. The former calculation
⫺1/ 2. The two Zeeman states are represented by the may alternatively be interpreted as 兩␤␤␣典 ⫽ 兩␤典 R 兩␤␣典.
following column vectors, This means we may use the already obtained represen-
tation of 兩␤␣典 in Eq. [68] and calculate 兩␤␤␣典 from the
1
兩␣典 ⫽ 0冉冊 [64]
direct product of a 2-dimensional vector (representing
兩␤典) with a 4-dimensional vector (representing 兩␤␣典).
The result is the following 8-dimensional vector:

0
冉冊

冢冣
兩␤典 ⫽ 1 [65] 0
0

冢冣
0 0
Coupled spin systems require larger dimensions of
the vectors representing the basis states. The spin
0
兩␤␤␣典 ⫽ 兩␤典 丢 兩␤␣典 ⫽ 1 冉冊丢 0 0
1 ⫽ 0 [69]
operators for a system of N I interacting spins-21 may be 0 0
represented in a Zeeman product basis (2, 3, 6, 34), 1
which is formed by taking direct products of the 0
Zeeman states of the individual spins.
The Zeeman product states form an orthonormal
兩m 1m 2 · · · m NI典 ⬅ 兩m 1典 丢 兩m 2典 丢 · · · 丢 兩m NI典 basis set:
[66]
具m⬘1m⬘2 · · · m⬘NI兩m 1m 2 · · ·m NI典
where R represents the direct product operation (3, 6, ⫽ ␦共m⬘1, m 1兲 䡠 ␦共m⬘2, m 2兲 · · · ␦共m⬘NI, m NI兲 [70]
34) and 兩m j 典 is either 兩␣典 or 兩␤典. It follows from
combinatorics that a system of N I spins-21 has 2 N I
different product basis states. Each Zeeman product state for a multiple-spin system
For instance, two spins-21 have four Zeeman prod- is simultaneously an eigenstate of the operators Îjz
uct states: {兩␣␣典, 兩␣␤典, 兩␤␣典, 兩␤␤典}. Each of these is (1 ⱕ j ⱕ N I ) for the individual spins, as well as of the
obtained as a 4-dimensional vector from a direct prod- operator for the z component of the total spin angular
uct operation: for example, state 兩␤␣典 is calculated as momentum Îz ⫽ Î1z ⫹ Î2z ⫹ . . . ⫹ ÎN Iz .
兩␤典 R 兩␣典. Note carefully that the order of the states in
Eq. [66] is important; for example, state 兩␣␤典 is dif- Î jz 兩m1 m2 · · · mN I典 ⫽ mj 兩m1 m2 · · · mN I典 [71]
ferent from 兩␤␣典. The direct product calculation is
carried out by multiplying each element 兩␤典 j of the
vector representation 兩␤典 by the entire vector 兩␣典: Î z 兩m1 m2 · · · mN I典 ⫽ 共m1 ⫹ m2 ⫹ · · ·
⫹ mN I兲兩m1 m2 · · · mN I典 [72]


兩␤典1 䡠 兩␣典
兩␤␣典⫽兩␤典 R 兩␣典⫽ 兩␤典 䡠 兩␣典
2
冊 [67]
Note that Îjz only operates on spin j and that all other
spins stay unaffected. In a multispin system the oper-
Using the explicit representations of Eqs. [64] and ator Î␨ represents the sum over all spins (within the
[65], the calculations are carried out according to system) having angular momentum component ␨
(where ␨ represents x, y, z, ⫹, or ⫺).

冉冊
冢 冣 冢冣
1 0
冉冊丢冉冊
0䡠 0
冘 Î
0 1 0 NI

冉冊
兩␤␣典 ⫽ 1 0 ⫽ ⫽ 1 [68]
1 Î ␨ ⫽ [73]
1䡠 0 0
j␨
j⫽1
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 131

For example, in a four-spin system the angular mo- 具␣␣␣兩Î 1z 1̂2 Î3x 兩␣␣␤典 ⫽ 具␣␣␣兩Î1z 兩␣␣␤典具␣␣␣兩1̂2 兩␣␣␤典
mentum raising operator Î⫹ corresponds to Î⫹ ⫽ Î⫹
1
⫹ Î⫹ ⫹ ⫹
2 ⫹ Î3 ⫹ Î4 and increases the z-angular momen-
⫻ 具␣␣␣兩Î3x 兩␣␣␤典
tum by one unit for all spins.
1 1
⫽ 䡠 1 䡠 具␣␣␣兩Î 3⫹ ⫹ Î3⫺ 兩␣␣␤典
2 2
Matrix Representations of Spin Operators
1
The matrix elements of spin operators can be calcu- ⫽ 具␣␣␣兩Î 3⫹ ␣␣␤典
4
lated in the Zeeman product basis using Eq. [71] and
the following expressions for the matrix elements of 1

the raising and lowering operators Î⫾ ⫹ ⫺
j (i.e., Îj or Îj ) 4
(4, 5, 28, 29):
where each underlined state indicates the “part” of the
⫾ bracket that is affected by the given operator. Likewise,
具m 1m 2 · · · m⬘j · · · m NI兩Î 兩m1 m2 · · · mj · · · mN I典
j
it may be shown that 具␤␤␤兩Î1zÎ3x兩␤␤␤典 ⫽ 0 and
⫽ 冑I共I ⫹ 1兲 ⫺ mj 共mj ⫾ 1兲␦共m⬘j , mj ⫾ 1兲 [74] 具␤␤␣兩Î1zÎ3x兩␤␤␤典 ⫽ ⫺1/4.
It is clear that this approach to calculate the ele-
Together with the following relations ments becomes very tedious for even moderately
large spin systems, because each operator is repre-
sented by a matrix of 2 N I ⫻ 2 N I dimensions. Instead,
1 it is much easier to construct matrix representations
Î jx ⫽ 共Î⫹ ⫹ Î⫺
j 兲 [75]
2 j using the direct product (3, 6, 34). The direct product
of two matrices A R B is effected by multiplying each
1 ⫹ element of the first matrix ( A jk ) by the entire matrix
Î jy ⫽ 共Î ⫺ Î⫺
j 兲 [76] B. In the case of spins-21, the operators are represented
2i j
by 2 ⫻ 2 matrices, and the resulting direct product
operators are of dimension 4 ⫻ 4.
this is sufficient to calculate the matrix representation
1 By exploiting the direct product, the matrix repre-
of any spin operator. For example, for a single spin-2,
sentation of the operator Î1z Î3x may be obtained with-
from Eqs. [71] and [74] the following matrices are
out extensive calculations as Î1z Î3x ⫽ Îz R 1̂ R Îx,
obtained (2– 6, 28, 29):
where the operators on the right-hand side correspond
to single spin-21 operators. In this case, the direct
Î z ⫽ 冉
1 1 0
2 0 ⫺1
冊 [77] product calculation involves forming an 8 ⫻ 8 matrix
from three 2 ⫻ 2 matrices. The explicit calculation is
given in Fig. 6, and further details of the construction

Î x ⫽ 冉 冊
1 0 1
2 1 0
[78]
of spin operators by direct products may be found in
Refs. (3, 6, 34). This “direct product calculation” is
the method of choice for numerical implementations.

冉 冊
Alternatively, another technique for fast evaluation of
1 0 1 matrix elements is described in Ref. (13).
Î y ⫽ [79]
2i ⫺1 0

SPATIAL TENSORS IN DIFFERENT


As an example of calculations of matrix elements
FRAMES
we construct some of the elements of the operator
Î1z Î3x, assuming a system of three coupled spins-21. Here we discuss how the spatial tensors of the NMR
This operator may be interpreted as as a product of interactions may be constructed and represented in
operators: Î1z Î3x ⬅ Î1z 䡠 1̂2 䡠 Î3x. Here 1̂ represents the various commonly used reference frames. We espe-
unity operator and the matrix representation for each cially focus on calculating [A ⌳ L
20 ] because it is needed
operator is of dimension 8 ⫻ 8. The index of the unity for obtaining the spin Hamiltonian (Eq. [60]).
operator has no significance other than for bookkeep-
ing reasons: it signifies that “spin number two” is
Principal Axis System
unaffected by the operator Î1z Î3x. For example, the
matrix element 具␣␣␣兩Î1z Î3x兩␣␣␤典 may be evaluated Each tensor has the simplest form in its principal axis
according to the following steps: system (PAS), denoted P. In this frame, the magni-
132 EDÉN

the asymmetry parameter is in the range 0 ⱕ ␩ ⱕ 1 (4,


5, 35 ). Note that other definitions of ␩ also exist in the
literature. If ␩ ⫽ 0, that is, if [␦ yy ] P ⫽ [␦ xx ] P , the
tensor is referred to as being axially symmetric (4, 5).
For all other values of ␩, the tensor is axially asym-
metric.
The principal values of the chemical shift tensor
may be extracted from the NMR spectrum. Due to
rapid molecular reorientations, the isotropic chemical
shift is the only directly observable parameter from
spectra of isotropic liquids because the anisotropic
parts average to zero. The spectrum corresponds to a
single sharp peak, as shown in Fig. 7(a). From the
spectrum of a static powder, on the other hand, the
principal values may be read off directly as indicated in
Fig. 7(b– d). These spectra are all for a constant shift
Figure 6 The recipe for constructing the matrix represen- anisotropy ␦aniso and different asymmetry parameters ␩.
tation of the operator Î1z Î3x for a system of three coupled Note how the spectral line shape depends on the value of
spins-21, starting from three 2 ⫻ 2 matrices of each individ- ␩. Also, a larger ␦aniso results in a broader spectrum.
ual spin. This is performed using the direct product A R B, Using numerical simulation programs, one may deter-
evaluated by multiplying each element A mn by the entire mine chemical shift tensor parameters by iterative fitting
matrix B. Here two consecutive direct products are required,
of calculated and experimental spectra.
which are evaluated from right to left as Î1z Î3x ⫽ Îz R (1̂ R
In the following we express the principal values in
Îx ). The first direct product creates a 4 ⫻ 4 matrix from two
2 ⫻ 2 matrices, and the second direct product operation frequency units, as well as the isotropic and anisotro-
gives the final matrix of dimension 8 ⫻ 8. pic shifts, by multiplying the various components
with the Larmor frequency:

tude of the tensor may be characterized by three ␻ iso ⫽ ␻0 ␦iso [84]


parameters, referred to as principal values, denoted
[A xx ] P , [A yy ] P , and [A zz ] P . For concreteness, we ␻ aniso ⫽ ␻0 ␦aniso [85]
continue the discussion with the chemical shift tensor
7
␦ CS, as an example.
Note that often the principal values are given in
In the deshielding convention (ppm), defined, for
shielding units ␴, leading to sign reversal of the
example, in Refs. (4, 6, 35), the chemical shift prin-
expressions for ␻iso and ␻aniso (1, 4, 35). From the
cipal values are denoted [␦ xx ] P , [␦ yy ] P , and [␦ zz ] P .
equations above, the zeroth and second rank irreduc-
From these it is convenient to introduce three new
ible tensors describing the chemical shift interaction
parameters,
in its principal axis system may be expressed as
1
␦ iso ⫽ 共关␦xx 兴P ⫹ 关␦yy 兴P ⫹ 关␦zz 兴P 兲
⫽ ⫺ 冑3 ␻iso
[80] 7 共0兲
3 ␻ CS ⫽ ␻00
CS
[86]

␦ aniso ⫽ 关␦zz 兴P ⫺ ␦iso [81] 7 共2兲 P


关␻ CS 兴 ⫽ 共关␻22 兴 , 关␻21
CS P
兴 , 关␻20
CS P
兴 , 关␻2⫺1
CS P CS P
兴 , 关␻2⫺2
CS P
兴 兲

␩⫽
关␦ yy兴 P ⫺ 关␦ xx兴 P
关␦ zz兴 P ⫺ ␦ iso
[82]
1

⫽ ␻aniso ⫺ ␩, 0,
2 冑 3
2
1

, 0, ⫺ ␩
2
[87]

where ␦iso is the isotropic chemical shift, correspond- 7


The rank two part ␻ CS(2)
is referred to as the chemical
ing to the average of the shift tensor over all orienta- shift anisotropy (CSA) tensor. Note that this is a
tions. ␦aniso and ␩ are called the anisotropy and the 7(2)
conflation of ␦ CS and B 0 and is strictly only a second
asymmetry parameter of the shift tensor, respectively. rank tensor as long as the magnetic field direction is
If the principal values are labeled such that constant, as discussed earlier.
Definitions similar to those given above apply to
兩 关␦ zz兴 P ⫺ ␦ iso兩 ⱖ 兩关␦xx兴P ⫺ ␦iso兩 ⱖ 兩关␦yy兴P ⫺ ␦iso兩 [83] any NMR interaction, provided that the relevant ten-
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 133

into a common reference system that is fixed on the


molecule; this molecular frame is denoted M and
illustrated in Fig. 8(a). This procedure is convenient
since all further transformations from the molecular
frame are identical for all interactions. The Euler
angles for transforming the second rank spatial tensor
of interaction ⌳ from the principal axis system to its
molecular frame are denoted ⍀ ⌳ PM . The transforma-
tion itself is carried out according to Eq. [52]:

7 7
关A ⌳共2兲 兴M ⫽ 关A⌳共2兲 兴P D̂共2兲 共⍀PM

兲 [89]

The tensor components in the molecular frame


⌳ M
[A 2m ] are related to those of the principal axis
system through Eq. [53]:


2
⌳ M ⌳ P 2 ⌳
关A 兴 ⫽
2m 关A 2m⬘ 兴 D m⬘m共⍀ PM 兲 [90]
m⬘⫽⫺2

Laboratory Frame
According to Eq. [60], only the m ⫽ 0 component of
a tensor expressed in the laboratory frame, is required
to construct the Hamiltonian. The following two sub-
sections give the general expressions for the spatial
tensor component A 20 in the laboratory frame, which
are appropriate for static and rotating solids, respec-
tively.
Figure 7 (a) The NMR spectrum of a chemical shift
tensor from a sample in solution, displaying a sharp peak at Static Solid. Assuming a single interaction, the an-

the isotropic chemical shift ␦iso. (b– d) Spectra of a chemical gles ⍀ PL transform the second rank tensor from its
shift tensor of a spin in a molecular fragment in a static principal axis system to the laboratory frame. In a
powder. The spectra are for (b) one axially symmetric shift powder containing randomly distributed crystallites,

tensor and (c, d) two axially asymmetric tensors. The posi- the angle ⍀ PL is specific for each crystal orientation.
tions of the principal values of the tensor are shown in each The expression for [A ⌳ L
20 ] in Eq. [60] is
case and they may be extracted from the shape of each
powder pattern. (b– d) The chemical shift anisotropy param-

冘 关A
eter ␦aniso is the same in (b– d). 2
⌳ L ⌳ P ⌳
关A 兴 ⫽
20 2m兴 D m0
2
共⌳ PL 兲 [91]
m⫽⫺2

sor elements A lm from Tables 2 and 3 are used. Thus,
for a general anisotropy parameter A aniso the tensor
We may inspect the dependence of the tensor on each
components are ⌳ ⌳ ⌳
of the individual Euler angles {␣ PL , ␤ PL , ␥ PL } by
7
冉 1
关A 共2兲 兴P ⫽ Aaniso ⫺ ␩, 0,
2 冑 3
2
1
, 0, ⫺ ␩
2
冊 [88]
using Eq. [49]

⌳ ⌳ ⌳ ⌳
2
D m0共⍀PL兲 ⫽ exp兵⫺im␣PL 其d m0
2
共␤PL 兲exp兵⫺i0␥PL 其 [92]
Molecular Frame
⌳ ⌳
⫽ d m0
2
共␤ PL 兲exp兵⫺im␣PL 其 [93]
If one needs to consider several NMR interactions
simultaneously in a calculation, it is common practice
to transform the tensor representing each interaction and rewriting Eq. [91] as
134 EDÉN


Table 3 Components A2m , ⴚ2 < m < 2, of Second Rank Spatial Tensors in Their Principal Axis System
Interaction (⌳) [A ⌳
20 ]
P ⌳
[A 2⫾1 ]P ⌳
[A 2⫾2 ]P

Chemical shift (CS) 冑 3


2
␦aniso
j
0 ⫺21 ␩␦aniso
j

Homonuclear dipolar (D) 公6 b jk 0 0


Heteronuclear dipolar (D) 公6 b IS
0 0

Homonuclear J 2␲ 冑 3
2
jk
J aniso 0 ⫺ 21 ␩J aniso
jk

2␲ 冑 3
Heteronuclear J 2
IS
J aniso 0 ⫺ 21 ␩J aniso
IS

First-order quadrupolar (Q) 冑 ␹ 3


2 Q 0 ⫺ 21 ␩␹ Q

The anisotropies A aniso of the spatial tensors (e.g., ␦aniso and J aniso
jk
) are defined as

A aniso ⫽ 关Azz 兴P ⫺ Aiso

with A iso given by


1
Aiso⫽3 ([Axx ]P ⫹ 关Ayy 兴P ⫹ 关Azz 兴P

The asymmetry parameter for each interaction is defined

␩ ⫽ 关A yy兴 P ⫺ 关A xx兴 P/共关A zz兴 P ⫺ A iso)

The dipolar coupling constants bjk and bIS are defined

b jk ⫽ ␮ 0␥ 2I ប/共4␲r jk3 兲 and bIS ⫽ ␮0 ␥I ␥S ប/共4␲rjk3 兲

with r jk and r IS being the distances between the two spins in the homonuclear and heteronuclear spin pair, respectively; ␮0 ⫽ 4␲ 䡠 107NC⫺2s2
7
is the permeability of vacuum and ប is Planck’s constant (h ⫽ 6.62608 䡠 10 ⫺34 Js) divided by 2␲. The quadrupolar tensor (Q (2) ) is defined
as the product of the electric field gradient tensor and the quadrupolar moment of the nucleus (eQ). With this definition, the magnitude of the
anisotropy of the quadrupolar interaction is given by the quadrupolar coupling constant ␹ Q ⫽ e 2 qQ/ប. The following units are employed for
the anisotropies: ␹ Q and b (rad s⫺1), ␦aniso (ppm), and J aniso (Hz).

ple, whereas the angle ⍀ ML is the same for all inter-


冘 关A
2
⌳ L ⌳ P 2 ⌳ ⌳
关A 20 兴 ⫽ 2m兴 d 共␤ PL
m0 兲exp兵⫺im␣PL 其 [94] actions of each crystal orientation. Therefore, the ten-
m⫽⫺2 sor for each interaction (⌳1, ⌳2, ⌳3, . . .) is usually
first transformed to the molecular frame using Eq.
Equation [94] shows that there is no dependence on [89]. The component [A ⌳ L
20 ] is related to the tensor
the Euler angle ␥ PL , which has implications when components in the molecular frame by
simulating NMR responses from powders. In a fol-
lowing publication, such orientational considerations
冘 关A
2
will be discussed in detail. ⌳ L
关A 20 兴 ⫽ ⌳ M
兴 D m0
2
共⍀ ML兲 [96]
2m
In the case of several interactions (⌳1, ⌳2, ⌳3, . . .), m⫽⫺2
it is useful to exploit transformations via the molec-
ular frame. Assuming the sequence of transformations
through the reference frames P 3 M 3 L [see Fig.
冘 关A
2
⌳ M 2
8(a) for a physical interpretation of these transforma- ⫽ 2m 兴 d 共␤ ML兲exp兵⫺im␣ML 其
m0 [97]
tions], the laboratory frame component [A ⌳ 20 ]
L
is m⫽⫺2

given by
where the last equality follows by using Eq. [49].


2
⌳ L ⌳ P 2 ⌳
关A 20 兴 ⫽ 关A 2m⬘ 兴 D m⬘m共⍀ PM 兲 D m0
2
共⍀ ML兲 [95] Rotating Solid. In the theoretical treatment of MAS
m,m⬘⫽⫺2 experiments, it is convenient to define a reference
frame R on the sample holder (rotor) such that the z
In a powder, the angle ⍀ ⌳ PM is specific for each axis of that “rotor frame” is along the sample rotation
interaction but common to all crystallites in the sam- axis [Fig. 8(b)]. Assume that the sample is spun at the
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 135

tory frame L. For MAS these are at a certain instant of


time t given by

⍀ RL共t兲 ⫽ 兵␣ RL共t兲, ␤ RL, ␥ RL其 ⫽ 兵⫺␻r t, ␪m , 0其 [99]

with the “magic angle” is given by ␪ m ⫽ arctan 公2.


In the case of sample rotation off the magic angle,
␤ RL takes an arbitrary value.
For practical calculations of rotating samples, the
transformation of a general second rank spatial tensor
from an arbitrary molecular frame to the laboratory
frame is performed by the following sequence of
rotations [depicted in Fig. 8(b)]:

7 ⍀MR 7 ⍀RL 7
关A ⌳共2兲 兴M O
¡ 关A⌳共2兲 兴R O
¡ 关A⌳共2兲 兴L [100]

Following the same calculations as in the static case,


the tensor in the laboratory frame is expressed

7 7
关A ⌳共2兲 兴L ⫽ 关A⌳共2兲 兴M D̂共2兲 共⍀MR 兲D̂共2兲 共⍀RL 共t兲兲 [101]

After carrying out the transformations and using Eq.


[49], the component [A ⌳ L
20 ] may be written


2
⌳ L ⌳ M 2
关A 20 兴 ⫽ 关A 2m⬘ 兴 D m⬘m共⍀ MR兲d m0
2
共␪ m兲exp兵im␻r t其
m⬘,m⫽⫺2

[102]
Figure 8 Visualization of different frames and transfor-
mations employed in solid-state NMR. Each principal axis and identified with a time-dependent frequency
system of two spatial tensors are shown (P V , P W ). Each of ␻ ⌳ (t) ⫽ [A ⌳ L
20 ] . Inspection of Eq. [102] reveals that
these is related to the molecular frame M by the Euler it is a Fourier series; it conforms to Eq. [25] with ␻mod
angles ⍀ PM
V
and ⍀ PM
W
, respectively. (a) The transformations
given by ␻ r ,
relevant for a calculation of a static sample. (b) Analogous
transformations for a calculation of a rotating sample. The
冘 冘
2 2
rotor frame R is chosen such that its z axis subtends the ⌳ M 2
关A 2m⬘ 兴 D m⬘m共⍀ MR兲d m0
2
共␪ m兲
angle ␤ RL with the static field direction. For magic-angle ␻ 共t兲 ⫽
⌳ m⫽⫺2 m⬘⫽⫺2
spinning, ␤ RL ⫽ arctan公2. The coordinate system M is
␻ ⌳共m)
related to the rotor frame by the angles ⍀ MR , and the final
transformation into the laboratory frame is given by the ⫻ exp兵im␻r t其 [103]

冘␻
time-dependent angles ⍀ RL (t) ⫽ {⫺␻ r t, ␪ m , 0}. 2
共m兲
⫽ ⌳ exp兵im␻r t其 [104]
m⫽⫺2

magic angle at a fixed angular frequency ␻ r , associ- and the five Fourier components ␻ ⌳
(m)
related to the
ated with a rotational period ⌳ M
tensor components [A 2m⬘ ] according to

␶ r ⫽ 2␲/␻ r [98]

2

␻ ⌳共m兲 ⫽ ⌳ M 2
关A 2m⬘ 兴 D m⬘m共⍀ MR兲d m0
2
共␪ m兲 [105]
corresponding to the time interval for the rotor to m⬘⫽⫺2

make a complete revolution. The sample rotation


gives rise to time-dependent Euler transformation an- The frequency ␻ ⌳ (t) is a real number, and the Fourier
gles ⍀ RL (t) relating the rotor frame R to the labora- components ␻ ⌳
(m)
and ␻ ⌳ (⫺m)
are related by
136 EDÉN

␻ ⌳共m兲 ⫽ 兵␻ ⌳共⫺m兲其* [106] NMR TIME-DOMAIN SIGNAL AND


SPECTRUM
This relationship is useful, for example, to speed up The remainder of the article discusses how the NMR
numerical calculations, because it follows that only signal in both the time and frequency domains is
␻⌳(m)
for m ⱖ 0 needs to be calculated explicitly by obtained as a result of the time evolution of the spin
Eq. [105]. This limits the calculation to only three out system. In this section a general formalism for calcu-
of five components, because the remaining two may lating the time-domain signal under any experimental
be obtained directly from Eq. [106]. Consequently, circumstance is outlined. The next section discusses
Eq. [104] may be written concepts of dynamically homogeneous and inhomo-
geneous Hamiltonians. Next, the form of the NMR
signal is derived for the case of a dynamically inho-
冘 Re兵␻
2
共0兲 共m兲 mogeneous Hamiltonian. Two specific classes of dy-
␻ ⌳共t兲 ⫽ ␻ ⫹ 2
⌳ ⌳ 其cos兵m␻r t其
namically inhomogeneous problems are examined in
m⫽1
more detail: experiments with a static sample and a
⫺ Im兵␻⌳共m兲 其sin兵m␻r t其 [107] rotating sample. Computer code for simulating NMR
time-domain signals and frequency-domain spectra
for these cases will be provided in the two following
Equations [106] and [107] are proven in Appendix A.
articles.
Equation [104] shows how MAS induces a time-
periodic modulation of the second rank spatial ten-
sors. The value of ␻ ⌳ (t 0 ) is equal to that following Density Operator
completion of a full rotational period: ␻ ⌳ (t 0 ) ⫽
␻ ⌳ (t 0 ⫹ ␶ r ). This may be verified by inserting the NMR experiments are normally performed on a very
large ensemble of nuclear spin systems, where the
two different time points into Eq. [104]. Using ␻ r ␶ r ⫽
spins may interact with each other within each sys-
2␲, the exponential factor may be expressed as
tem, but the various ensemble members (i.e., spin
exp{im␻ r (t 0 ⫹ ␶ r )} ⫽ exp{im␻ r t 0 }exp{2im␲} ⫽
systems) are independent of each other. In such cases,
exp{im␻ r t 0 }. This demonstrates that the spatial part
the density operator formalism is applicable. This is a
of the Hamiltonian is periodic in time. statistical approach, having the advantage that one
obtains a density operator ␳ˆ , representing the entire
ensemble state, without the need to consider the state
Rotating Frame of each individual spin system in the very large en-
semble. Further information about the density opera-
Although we have thus far stressed the laboratory tor formalism may be found in Refs. (2– 6, 29, 34).
frame as the “final” reference frame, in practice an The density operator is represented by the density
additional transformation of the Hamiltonian is nor- matrix in a suitable basis set. For example, for an
mally employed: one to the rotating frame (1– 6 ). ensemble of isolated spins-21, the density matrix may
This involves transforming the spin parts of the Ham- in the Zeeman basis be written
iltonian, and it must be distinguished from the rotor
frame R of the previous section. Employing the rotat-
ing frame means observing the NMR experiment from 冉
具␣兩␳ˆ 兩␣典 具␣兩␳ˆ 兩␤典
␳ˆ ⫽ 具␤兩␳ˆ 兩␣典 具␤兩␳ˆ 兩␤典 冊 [108]
a time-dependent coordinate system, where the trans-
verse plane rotates around its z axis (which coincides
Each element of the density matrix represents the spin
with that of the laboratory frame) at a constant rate
ensemble in various states: the diagonal elements
␻ref, close to the Larmor frequency of the spins.
correspond to fractional populations of the states,
The resulting rotating frame spin Hamiltonian in- whereas the off-diagonal elements represent coher-
cludes all interactions of the local magnetic fields, ences. We refer to Refs. (2– 6) for the exact meanings
such as the chemical shifts and dipolar couplings, but of these terms. In the case of isolated spins-2, the
1

excludes the Zeeman interaction. The rotating frame elements 具␣兩␳ˆ 兩␣典 and 具␤兩␳ˆ 兩␤典 give the fractional pop-
Hamiltonian may be obtained from the laboratory ulations of the states 兩␣典 and 兩␤典, respectively. If the
frame Hamiltonian after subtracting the term ␻refÎz magnitudes of these elements are different there is a
from the latter. The explicit transformations are given net longitudinal spin polarization (and magnetization)
in Refs. (2– 6, 35). along the static field direction. The two off-diagonal
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 137

elements 具␣兩␳ˆ 兩␤典 and 具␤兩␳ˆ 兩␣典 correspond to ⫺1 and This expression may be derived by multiplying the
⫹1 quantum coherences (⫺1QC and ⫹1QC), respec- density operator from both sides by the unity operator,
tively. The presence of single-quantum coherences in 1̂␳ˆ 1̂, and then inserting the closure relation (29)
the ensemble is equivalent to a net spin polarization
(and magnetization) in the transverse plane of the
冘 兩 j典具 j兩 ⫽ 1̂

rotating frame. The spin density operator may be [112]
expressed as a linear combination of spin operators of j⫽1
the system. Net longitudinal magnetization corre-
sponds to a density operator proportional to Îz ,
which states that the sum over all projection operators
whereas transverse magnetization corresponds to ␳ˆ
兩 j典具 j兩 (2– 6, 29, 34) equals the unity operator. Then
being proportional to a combination of Îx and Îy , or
we obtain the expression
equivalently, a linear combination of Î⫹ and Î⫺. (See
Eqs. [75] and [76] for the relationship between the

冘 兩 j典具 j兩 䡠 ␳ˆ 䡠 兩k典具k兩

various operators.)
The density operator of an ensemble of isolated ␳ˆ ⫽ [113]
spins-21 may be written using the Dirac formalism as j,k⫽1

Equation [111] follows from interpreting the product



兩 j典具 j兩 䡠 ␳ˆ 䡠 兩k典具k兩 as 兩 j典 䡠 具 j兩␳ˆ 兩k典 䡠 具k兩 and then using
␳ˆ ⫽ 具 j兩␳ˆ 兩k典 䡠 兩 j典具k兩 [109] that 具 j兩␳ˆ 兩k典 as a complex number that may be rear-
j,k⫽␣,␤
ranged within the product. (Similar calculations were
carried out in an earlier section.)
⫽ 具␣兩␳ˆ 兩␣典 䡠 兩␣典具␣兩 ⫹ 具␣兩␳ˆ 兩␤典 䡠 兩␣典具␤兩 Note that the basis employed in Eq. [111] may not
necessarily be the Zeeman product basis. Formally,
⫹具␤兩␳ˆ 兩␣典 䡠 兩␤典具␣兩 ⫹ 具␤兩␳ˆ 兩␤典 䡠 兩␤典具␤兩 [110] the interpretation of the diagonal and off-diagonal
elements in terms of populations and coherences is
The explicit matrix representation of Eq. [110] corre- similar, regardless of the choice of basis. However,
sponds to Eq. [108]. The operators 兩␣典具␤兩 and 兩␤典具␣兩 unless the Zeeman basis is employed, the physical
correspond to ⫹1QC and ⫺1QC, respectively. For meanings of the matrix elements (e.g., in terms of spin
example, it may be verified that the operator 兩␤典具␣兩 polarizations and magnetizations) differ from the
corresponds to Î⫺ by letting it operate on each of the usual interpretations. In the following, we will fre-
Zeeman basis states 兩␣典 and 具␤兩. We get quently express the density operator in the eigenbasis
of the Hamiltonian, because this is well suited for
describing spin dynamics (10 –13, 15, 17, 18).
共兩␤典具␣兩兲 䡠 兩␣典 ⫽ 兩␤典 䡠 具␣兩␣典 ⫽ 兩␤典
The starting point for any NMR experiment is the
spin ensemble at thermal equilibrium in a strong mag-
and netic field B ᠬ . Over the ensemble, there is a net polar-
ization pointing along the static magnetic field (i.e., a
共兩␤典具␣兩兲 䡠 兩␤典 ⫽ 兩␤典 䡠 具␣兩␤典 ⫽ 0 net longitudinal polarization). As pointed out earlier,
this corresponds to a density operator being propor-
tional to Îz (2– 6, 34),
as expected from the operator Î⫺: it converts state 兩␣典
into state 具␤兩. Likewise, it may be demonstrated that
兩␣典具␤兩 ⫽ Î⫹. ␳ˆ eq ⬀ 共1̂ ⫹ ␤I Îz 兲 [114]
In the general case of an ensemble consisting of
multiple-spin systems, with the spin operators repre- where ␤ I is the Boltzmann factor, given by
sented in an arbitrary basis spanned by ᏺ states {兩1典,
兩2典, . . . , 兩ᏺ典}, the density operator may formally be ⫺ប␥I B0
expressed in the Dirac formalism as the following ␤I ⫽ [115]
kB T
sum:
Here T is the temperature (K) and k B ⫽ 1.381 䡠
10 ⫺23 JK⫺1 is the Boltzmann constant. The unity
冘 具 j兩␳ˆ 兩k典 䡠 兩 j典具k兩

␳ˆ ⫽ [111] operator and the Boltzmann factor will not affect any
j,k⫽1 of the results in our treatment and will be dropped.
138 EDÉN

The thermal equilibrium density operator used in the


following is then given by

␳ˆ eq ⫽ Îz [116]

Time Evolution of Density Operator


Assume the spin ensemble is initially at t ⫽ t a in a
certain state represented by ␳ˆ (t a ) and that the total
spin Hamiltonian (including the contributions from all
interactions) does not commute with ␳ˆ (t a ). Then the
Hamiltonian will cause a change of the spin ensemble
state, which mathematically translates into a change
of the density operator; it becomes time dependent:


␳ˆ 共t a兲 ¡ ␳ˆ 共t兲 [117]

Figure 9 A propagator Û(␶ ca ), which is schematically


The central question when dealing with quantum dy-
depicted as the shaded box at the bottom of the figure, is to
namics in NMR is, given that we know the density be determined over the time interval ␶ ca ⫽ t c ⫺ t a . This
operator at the time point t a , how do we find its propagator may be formed by first dividing the interval ␶ ca
expression at a later time point t b ? The solution to the into two smaller segments, ␶ ba ⫽ t b ⫺ t a and ␶ cb ⫽ t c ⫺
problem is the operator Û(t b , t a ), called the propa- t b , and then calculating Û(␶ ca ) as the time-ordered product
gator, which transforms the density operator accord- Û(␶ ca ) ⫽ (Û␶ cb ) Û(␶ ba ) from the propagators over the
ing to the following “sandwich formula” (2– 6, 29): smaller time segments. Note the order of the operators in the
product. If the expression for ␳ˆ (t a ) is known, each of these
␳ˆ 共t b兲 ⫽ Û共tb , ta 兲␳ˆ 共ta 兲Û共tb , ta 兲† [118] propagators may then be used to determine the density
operators ␳ˆ (t b ) and ␳ˆ (t c ) according to the operator sand-
wich formula, Eq. [118].
Note that the time points are written such that later
times appear to the left, for example, the propagator
transforming ␳ˆ (t a ) into ␳ˆ (t b ) (with t b ⬎ t a ) is written
Û(t b , t a ). Properties of Propagators
The propagator is a unitary operator (2– 6, 29), In this section we discuss a few important properties
from which follows that its inverse may be obtained of propagators, which will prove very useful in nu-
by the adjoint operation (see Eq. [4]): merical implementations.
Assume that the Schrödinger equation is integrated
Û共tb , ta 兲⫺1 ⫽ Û共tb , ta 兲† [119] over the interval ␶ ca ⫽ t c ⫺ t a , with t c ⬎ t a as
illustrated in Fig. 9. By definition, the solution is the
The propagator is formally related to the Hamilto- propagator Û(t c , t a ) ⬅ Û(␶ ca ), which transforms
nian through the Schrödinger equation (2– 6, 29) ␳ˆ (t a ) into ␳ˆ (t c ) through the operator sandwich, Eq.
[118]. At the moment, we do not consider how the
d propagator is determined. Here we introduce the
Û共t, ta 兲 ⫽ ⫺iĤ共t兲Û共t, ta 兲 [120]
dt shorthand notation Û(␶ ca ): in the following, we will
use the two notations Û(t c , t a ) and U(␶ ca ) inter-
This differential equation needs to first be solved for changeably.
an expression of Û(t b , t a ), which subsequently can be If a time point t b is picked within the interval ␶ ca ,
used in Eq. [118]. However, the Schrödinger equation we may solve the Shrödinger equation separately over
generally has no exact solution. This is one reason each of the two time segments ␶ ba and ␶ cb of Fig. 9.
why computer programs that numerically carry out This results in Û(t b , t a ) ⬅ Û(␶ ba ) [transforming ␳ˆ (t a )
the integration are useful. On the other hand, in the 3 ␳ˆ (t b )] and Û(t c , t b ) ⬅ Û(␶ cb ) [transforming ␳ˆ (t b )
case of dynamically inhomogeneous Hamiltonians, 3 ␳ˆ (t c )]. The following relationship holds between
Eq. [120] may be solved analytically. We will dem- the three operators, Û(t c , t a ), Û(t b , t a ), and Û(t c , t b )
onstrate this in a later section. (29):
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 139

Û共tc , ta 兲 ⫽ Û共␶cb 兲Û共␶ba 兲 [121] in 3-dimensional space, that is, transforming an object
by a sequence of rotation operators R̂(⍀ FF⬘ ). In the
⫽ Û共tc , tb 兲Û共tb , ta 兲 [122] former case, the propagator Û(t c , t a ) may be decom-
posed into a product of propagators over subsegments
The order of the propagators in the product Û(tc, tb) of ␶ ca , according to Eq. [122]. In the latter case, a net
Û(tb, t a ) is important and may not be changed, be- rotation operator may be decomposed into a product
cause the operators generally do not commute. A of rotation operators (each effecting a “simpler rota-
product of the form of Eq. [122] is said to be time tion”) as illustrated by Eq. [39].
ordered (1–5): propagators involving later time Another important property of propagators is that
points appear to the left to those involving earlier time Û(t a , t b ) is the inverse of Û(t b , t a ), that is,
points.
We motivate Eq. [122] as follows: if we are to Û共ta , tb 兲 ⫽ 兵Û共tb , ta 兲其⫺1 ⫽ Û共tb , ta 兲† [127]
propagate the density operator from t ⫽ t a to t ⫽ t c ,
we may either do the transformation in one step using Note the ordering of time points. If t b ⬎ t a , the
the operator of the left-hand side of Eq. [122] accord- meaning of Û(t a , t b ) is the operator that transforms
ing to the density operator backward in time: ␳ˆ (t b ) 3 ␳ˆ (t a ).
Equation [127] may be proven as follows: assume that
␳ˆ 共t c兲 ⫽ Û共tc , ta 兲␳ˆ 共ta 兲Û共tc , ta 兲† [123] we first transform ␳ˆ (t a ) into ␳ˆ (t b ) according to Eq.
[118]. Next, we perform the reverse transformation
or through two successive transformations using the ␳ˆ (t b ) 3 ␳ˆ (t a ) according to
right-hand side of Eq. [122]. We verify that the sand-
wich ␳ˆ 共t a兲 ⫽ Û共ta , tb 兲␳ˆ 共tb 兲Û共ta , tb 兲† [128]

兵Û共tc , tb 兲Û共tb , ta 兲其␳ˆ 共ta 兲兵Û共tc , tb 兲Û共tb , ta 兲其† [124] By combining both transformations sequentially, we
get the following result:
is indeed equal to the density operator ␳ˆ (t c ). Using
the fact that the adjoint of a product of two unitary
␳ˆ 共t a兲 ⫽ Û共ta , tb 兲 䡠 ␳ˆ 共tb 兲 䡠 Û共ta , tb 兲† [129]
operators is equal to the reversed product involving
the adjoint of each operator (6, 29), that is, Û共t b , t a 兲␳ˆ 共t a 兲Û共t b , t a 兲 †

⫽ 兵Û共ta , tb 兲Û共tb , ta 兲其␳ˆ 共ta 兲兵Û共ta , tb 兲Û共tb , ta 兲其†


兵ÂB̂其† ⫽ B̂† † [125]
[130]
it follows that the rightmost factor of Eq. [124] cor-
responds to {Û(t c , t b )Û(t b , t a )} † ⫽ Û(t b , t a ) † Û(t c , Because the left-hand and right-hand sides of Eq.
t b ) † . Hence, we may write [130] must be equal, it follows that each of the oper-
ator products within braces must be equal to the unity
Û共tc , tb 兲Û共tb , ta 兲 䡠 ␳ˆ 共ta 兲 䡠 Û共tb , ta 兲† Û共tc , tb 兲† operator,

⫽ Û共tc , tb 兲 䡠 Û共tb , ta 兲␳ˆ 共ta 兲Û共tb , ta 兲† 䡠 Û共tc , tb 兲† [126] Û共ta , tb 兲Û共tb , ta 兲 ⫽ 1̂ [131]
␳ˆ 共t b 兲
which proves Eq. [127]. The procedure employed to
Inserting Eq. [118] into Eq. [126] gives Û(t c , demonstrate Eq. [127] also has an analogy to rotations
t b )␳ˆ (t b )Û(t c , t b ) † , which, by definition, equals ␳ˆ (t c ). in space: an object stays unaffected if it is first rotated
Hence, the right-hand side of Eq. [122] effects the by an angle ␪ around a given axis and subsequently
transformation ␳ˆ (t a ) 3 ␳ˆ (t c ). rotated by the angle ⫺␪ around the same axis.
In general, if the density operator is to be propa-
gated over a given time interval ␶ ca , the result is the
NMR Signal
same, regardless of whether the propagator Û(t c , t a )
is used directly or a series of “small-step” transfor- The previous sections explained how the ensemble of
mations (involving propagators over smaller time seg- spin systems are represented by a density operator and
ments within ␶ ca ) are used successively. how the Hamiltonian causes the density operator to
Time propagation of the density operator, that is, evolve in time through the propagator. We now con-
transforming ␳ˆ (t a ) 3 ␳ˆ (t c ) in successive steps, is tinue to discuss the relationship between the density
analogous to a series of geometrical transformations operator and the NMR time-domain signal.
140 EDÉN

The acquired NMR time-domain signal s(t) corre-


sponds to the expectation value 具Q̂典 of an “observable
operator” Q̂. In principle, Q̂ may be any Hermitian
spin operator, because we may only directly detect
quantized systems through Hermitian operators (29).
The time-domain signal s(t) ⫽ 具Q̂典(t) may be calcu-
lated from the trace of the product between the density
operator and the observable operator:

s共t兲 ⫽ 具Q̂典共t兲 ⫽ Tr兵␳ˆ 共t兲Q̂其 [132] Figure 10 A radio frequency pulse scheme for a simple
NMR experiment generating a detectable NMR time-do-
main signal. The spin ensemble is initially at thermal equi-
The trace of an operator corresponds to the sum of its
librium, corresponding to the density operator ␳ˆ eq ⫽ Îz . A
diagonal elements (29). Using an arbitrary basis set,
␲/2 pulse is applied along the y axis of the rotating frame,
the time-domain signal may be alternatively ex- creating observable single-quantum coherences (“transverse
pressed in either of the following forms: magnetization”). The NMR signal acquisition starts right
after the pulse and defines t ⫽ 0. At this time point, the
density operator is given by ␳ˆ (0) ⫽ Îx ⫽ Î⫹ ⫹ Î⫺, where the
冘 具 j兩␳ˆ Q̂兩 j典

presence of ⫹1QC and ⫺1QC is signified by the operators
s共t兲 ⫽ [133]
Î⫹ and Î⫺, respectively. The subsequent evolution of the
j⫽1
⫺1Q coherences generates the NMR time-domain signal
(here illustrated by a damped oscillating function), as dis-
冘 具 j兩␳ˆ 兩k典具k兩Q̂兩 j典

cussed in the text. The figure is not drawn to scale: the
s共t兲 ⫽ [134]
duration of the pulse is on the order of microseconds
j,k⫽1
whereas the signal acquisition interval is normally in the
range of tens or hundreds of milliseconds.
Using Eq. [133], the signal is calculated by taking the
trace of the product ␳ˆ Q̂, while Eq. [134] implements
the trace as a pairwise multiplication of elements of ␳ˆ 共t 0兲 ⫽ Î x [135]
the two operators. The latter expression follows after
inserting the closure relation into Eq. [133]; a similar Because Îx has off-diagonal elements (as may be
calculation was performed earlier. verified from its matrix representation, Eq. [78]) it
Modern NMR spectrometers employ quadrature represents a spin ensemble state of single-quantum
detection, which results in detection of ⫺1QC (4 – 6, coherences. As discussed above, single-quantum co-
35). In practice, this is performed by simultaneously herences may be interpreted as transverse magnetiza-
recording the expectation values of the two Hermitian tion, which may be detected by the NMR spectrom-
operators 具Îx 典(t) and 具Îy 典(t). Next, they are combined eter. The NMR time-domain signal generated by the
using Eqs. [75] and [76] to obtain the expectation subsequent spin evolution under the Hamiltonian is
value 具Î⫹ 典(t). Therefore, Î⫹ corresponds to the de- usually referred to as a free-induction decay (2– 6).
tection of ⫺1QC, despite the fact that formally, this The time point at the start of NMR signal acquisition
operator is not Hermitian (as may be seen from its is denoted t 0 , and it is furthermore defined to be equal
matrix representation). We refer to Refs. (4 – 6, 35) to zero (2– 6). Figure 10 depicts the RF scheme of the
for more complete accounts of quadrature detection. pulse and the corresponding changes of the density
operator. Assuming that the time-domain signal is
Single Pulse Experiment. Assume an RF pulse of acquired by this simple one-pulse experiment, its cor-
flip angle ␲/2 and phase ␲/2 [denoted (␲/ 2) y , see responding NMR frequency-domain spectrum S(␻) is
Refs. (2– 6)] is applied to an ensemble of nuclear obtained by Fourier transforming s(t) using Eq. [28],
spins at thermal equilibrium in a strong magnetic as discussed previously.
field. Before the pulse is initiated, there is a net spin Equations [118], [120], and [132] are the funda-
polarization pointing along the static field. The pulse mental expressions for calculating the NMR time-
rotates this polarization around the rotating frame y domain signal from any experiment. However, these
axis by the angle ␲/2, leaving the net polarization equations are very general and in their present form
pointing along the rotating frame x axis right after the do not provide much insight into the dynamics. The
pulse. This corresponds to the following density op- following sections discuss how they are applied to
erator so-called “dynamically inhomogeneous problems”
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 141

and how the properties of the resulting NMR signals Maricq and Waugh (27) introduced the nomencla-
in the time and frequency domains are derived. The ture dynamically inhomogeneous Hamiltonian when it
resulting equations will be implemented numerically is self-commuting and dynamically homogeneous
in my next article. when it is not self-commuting. In the former case,
It should be kept in mind that the spin Hamiltonian, [Ĥ(t 1 ), Ĥ(t 2 )] ⫽ 0. Physically, the spin evolution
density matrix, and NMR signal are not only depen- under a sum of mutually commuting Hamiltonians
dent on time but also on the spatial orientation of the (i.e., the dynamically inhomogeneous case) is the
molecule: For brevity, this dependence has not been superposition of the evolution under each interaction
indicated explicitly, but it will be included in a fol- taken in isolation: the various interactions “add up.”
lowing article. Another factor that needed to be taken However, if these Hamiltonians do not commute (i.e.,
into account in a numerical simulation is relaxation of the dynamically homogeneous case), the net evolution
the nuclear spins: we do not include this explicitly,
is more complicated because of interference between
but the damping of the spin coherences (T2 or spin–
the various terms. An important point is that the total
spin relaxation) is accounted for phenomenologically,
Hamiltonian is dynamically homogeneous, even if all
as discussed later.
individual Hamiltonians commute with each other
except at least one (27).
Consequently, it is necessary to identify each ex-
DYNAMICALLY HOMOGENEOUS AND perimental situation that is to be simulated numeri-
INHOMOGENEOUS HAMILTONIANS cally and determine whether its Hamiltonian is homo-
geneous or inhomogeneous. This depends both on the
The crucial step in NMR simulations is the numerical “internal” interactions within the spin system and on
integration of the Shrödinger equation (Eq. [120]). the experimental conditions (MAS and RF fields).
Therefore, perhaps the most important factor dictating The Hamiltonians for experiments with rotating sol-
both the complexity of computer implementation (i.e., ids are dynamically homogeneous in general. How-
the code) and the time span when executing the sim- ever, there are some important exceptions. Examples
ulation is the nature of the Hamiltonian under which of dynamically inhomogeneous cases in rotating sol-
the spins evolve. ids are isolated nuclear spins subject to chemical shift
If the Hamiltonian does not commute with itself anisotropy and/or first-order quadrupolar interactions.
when evaluated at two different time points, that is, if Heteronuclear spin systems also evolve under dynam-
ically inhomogeneously Hamiltonians, as long as the
关Ĥ共t1 兲, Ĥ共t2 兲兴 ⫽ Ĥ共t1 兲Ĥ共t2 兲 ⫺ Ĥ共t2 兲Ĥ共t1 兲 ⫽ 0 [136] spins are not subject to second-order quadrupolar
interactions or homonuclear couplings: the hetero-
nuclear dipolar or J-coupling Hamiltonians commute
the Schrödinger equation generally has no simple with the chemical shifts or first-order quadrupolar
analytical solution. The following article will outline
Hamiltonians, as can be verified using the expressions
how Û(t, t 0 ) may be approximated numerically in
of the spin operators given in Table 2. Furthermore, as
such cases. This article is primarily confined to deal-
discussed in Ref. (4), special cases exist in rotating
ing with somewhat simpler problems: those for which
solids where the spin operators of the interactions do
the Hamiltonian is either time-independent or self-
commuting and periodic in time. not commute with each other, but which are never-
If the Hamiltonian is time independent (Ĥ ⫽ con- theless dynamically inhomogeneous because all spa-
stant), the propagator, which is obtained as the solu- tial tensor orientations coincide. One example is a
tion to Eq. [120], is given by the simple expression linear chain of homonuclear dipolar-coupled spins
with identical chemical shift tensors.
Note, however, that all the cases mentioned
Û共t, t0 兲 ⫽ exp兵⫺i共t ⫺ t0 兲Ĥ其 [137] above are dynamically inhomogeneous only in the
absence of RF fields: the rotating frame RF Ham-
This allows the obtaining of the density operator, and iltonian during a pulse (applied exactly on reso-
hence the time signal, at any point in time by the nance) of constant amplitude ␻RF and phase ␾RF
evaluation of an exponential operator, which can be may be written (1– 6 )
easily done using the results presented earlier. An
experiment with a static sample is a typical example
of such a case. Ĥ RF ⫽ ␻RF共Îx cos兵␾RF其 ⫹ Îy sin兵␾RF其兲 [138]
142 EDÉN

冦冕 冧
It may be verified that ĤRF does not commute with the t
internal spin interactions listed in Table 2 and hence
will lead to homogeneous spin dynamics. Û共t, t0 兲 ⫽ exp ⫺i dt⬘Ĥ共t⬘兲 [141]

t0

冦 冘冕 冧
t
NMR TIME SIGNAL AND SPECTRUM

FROM DYNAMICALLY INHOMOGENEOUS
HAMILTONIAN ⫽ exp ⫺i ␻u 共t⬘兲dt⬘兩u典具u兩 [142]
u⫽1
t0
General Formalism
Here we derive the general expression for the NMR where the last equality follows by inserting Eq. [140].
time-domain signal (Eq. [132]) from spin systems Equation [142] shows that the propagator corresponds
evolving under dynamically inhomogeneous Hamil- to a sum over complex exponentials of integrated
tonians. Similar treatments may be found, for exam- Hamiltonian eigenvalues, each multiplied with the
ple, in Refs. (2, 4, 36). For most equations, we keep projection operator 兩u典具u兩. We define the real-valued
the arbitrary initial time point t 0 for completeness function ⌽ u (t, t 0 ) as the integrated eigenvalue ␻ u (t)
(e.g., in Eqs. [118] and [120]), but we always assume over the time interval
that t 0 ⫽ 0 (corresponding to the start of acquisition
in an NMR experiment) when evaluating the NMR
t


signal. We employ a rotating frame in the analysis and
assume the absence of RF fields during the acquisi-
⌽ u共t, t 0兲 ⫽ dt⬘␻ u共t⬘兲 [143]
tion.
t0
Dynamically Inhomogeneous Hamiltonians. Dy-
namically inhomogeneous Hamiltonians have, in gen-
The symbol ⌽ u (t, t 0 ) is called the dynamic phase (4,
eral, time-dependent eigenvalues but time-indepen-
36 –38) accumulated by the eigenstate 兩u典 over the
dent eigenstates (eigenvectors) (27). Consequently, if
time interval t 0 ⱕ t⬘ ⱕ t. The dynamic phases
the Hamiltonian is diagonalized, a set of ᏺ eigenval-
involve products of Hamiltonian eigenvalues (rads⫺1)
ues ␻ u and ᏺ orthonormal eigenstates 兩u典 is obtained,
and time (s); hence, ⌽ u (t, t 0 ) is in rad units, which
where each state obeys the following eigenequation:
means it may be interpreted as an “angle.” We elab-
orate on this physical interpretation further later when
Ĥ共t兲兩u典 ⫽ ␻u 共t兲兩u典 [139] considering the time evolution of the density operator
and the form of the time-domain signal. From the
properties of the integral it follows that the dynamic
We reserve the label u (and when needed v, r, and s)
phase ⌽ u (t 0 , t) is related to ⌽ u (t, t 0 ) by sign reversal:
for indices representing eigenstates and eigenvalues
of the Hamiltonian or its corresponding propagator.
As for any operator represented in its eigenbasis (Eq. t0 t

[7]), the Hamiltonian may be expanded as products of


an eigenvalue ␻ u (t) and a projection operator 兩u典具u兩
onto the corresponding eigenstate 兩u典 (2, 4, 36)
⌽ u共t 0, t兲 ⬅ 冕 冕
dt⬘␻ u共t⬘兲 ⫽ ⫺ dt⬘␻u 共t⬘兲

t t0

⬅ ⫺⌽u 共t, t0 兲 [144]


冘 ␻ 共t兲兩u典具u兩

Ĥ共t兲 ⫽ u [140]
u⫽1 By Taylor expansion of the exponential operator
exp{iÂ} (Eq. [13]) one can show that it commutes
Propagator and Dynamic Phase. Next we seek an with Â, and consequently the two operators share the
expression for the propagator. If the Hamiltonian same eigenbasis (29). It then follows that the propa-
commutes with itself at all time points between t 0 and gator is diagonal in the Hamiltonian eigenbasis, and
t (as in the present case), the formal solution to Eq. Eq. [142] may be expressed in terms of dynamic
[120] is obtained as the exponential of the integral of phases ⌽ u (t, t 0 ) and projection operators 兩u典具u兩 as
the Hamiltonian over the time interval t 0 ⱕ t⬘ ⱕ t, (4, 36)
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 143

⌽ uv共t, t 0兲 ⫽ ⌽ v共t, t 0兲 ⫺ ⌽ u共t, t 0兲



ᏺ [152]
Û共t, t0 兲 ⫽ exp兵⫺i⌽u 共t, t0 兲其兩u典具u兩 [145]
u⫽1
Note how the difference is formed: the phase corre-
sponding to the leftmost index (u) is subtracted from
which is a relatively simple expression. Equation that to the right (v). In the following, we employ
[145] is proven in Appendix B. The inverse propaga-
similar indexing to denote differences between other
tor Û(t, t 0 ) † , is given by
entities.
Equation [151] shows that the density operator

冘 exp兵i⌽ 共t, t 兲其兩u典具u兩



remains proportional to the operator 兩u典具v兩 at all time
Û共t, t0 兲† ⫽ u 0 [146] points. All changes are accommodated in the phase
u⫽1
factor exp{i⌽ uv (t, t 0 )}. The spin dynamics under a
dynamically inhomogeneous Hamiltonian is remark-
Hence, the operators Û(t, t 0 ) † and Û(t, t 0 ) only differ ably simple; if the spin ensemble is initially in a
by the sign of their exponents. certain state, it stays there at all times. We stress that
Eq. [151] is a direct consequence of having time-
Time Evolution of Density Operator. We may get
independent Hamiltonian eigenstates. All manipula-
more insight into the role of the dynamic phase in spin
tions leading to Eq. [151] implicitly exploited that the
dynamics by examining the time evolution of the
eigenstates of the Hamiltonian and propagator are
density operator that the propagator of Eq. [145] gen-
independent of time.
erates: assume that the density operator is initially
A homogeneous Hamiltonian, on the other hand,
proportional to an arbitrary operator 兩u典具v兩: ␳ˆ (t 0 ) ⫽
兩u典具v兩. This operator is associated with a coherence has time-dependent eigenstates: the corresponding
between the Hamiltonian eigenstates 兩u典 and 兩v典. propagator (i.e., the solution of Eq. [120]) generally
The density operator at a later time point t is has no closed analytical form. Assume that the density
obtained by combining Eqs. [145] and [118] as fol- operator is initially proportional to 兩u典具v兩 (represent-
lows: ing a certain coherence in the spin ensemble) and
subject to a dynamically homogeneous Hamiltonian.
At a later time point, this coherence will generally be
␳ˆ 共t兲 ⫽ Û共t, t0 兲␳ˆ 共t0 兲Û共t, t0 兲† [147]
transferred into other coherences and populations, de-

冘 exp兵⫺i⌽ 共t, t 兲其兩r典具r兩


ᏺ pending on the exact form of the Hamiltonian. The
⫽ r 0
density operator at a given time point may always be
r,s⫽1 expressed as a linear combination of its eigenstates.
Û共t, t 0 兲
However, these are time dependent for dynamically
homogeneous Hamiltonians.
⫻ 兩u典具v兩 䡠 exp兵i⌽s共t, t0兲其兩s典具s兩 [148]
␳ˆ 共t0 兲 Û共t, t0 兲† NMR Time-Domain Signal. In this section we derive
the general expression for the NMR time-domain

冘 exp兵⫺i⌽ 共t, t 兲其exp兵i⌽ 共t, t 兲其兩r典



signal from a dynamically inhomogeneous Hamilto-
⫽ r 0 s 0 nian (Eq. [159]) for t 0 ⫽ 0.
r,s⫽1
The time-domain signal is calculated as the trace of
⫻ 具r兩u典 䡠 具v兩s典 䡠 具s兩 [149] the product of the density and observable operators
(Eq. [133]). The trace of an operator (or products of
Because the Hamiltonian eigenstates are orthonormal, operators) is independent of the choice of basis set for
we may use Eq. [2]: 具r兩u典 ⫽ ␦ ru and 具v兩s典 ⫽ ␦ vs . The the matrix representation (3, 4, 29): we employ the
sum then reduces to the single term eigenbasis of the Hamiltonian, leading to

␳ˆ 共t兲 ⫽ exp兵⫺i⌽u 共t, t0 兲其exp兵i⌽v 共t, t0 兲其 䡠 兩u典具v兩 [150]


冘 具u兩␳ˆ 共t兲Q̂兩u典

⫽ exp兵i⌽uv 共t, t0 兲其兩u典具v兩 [151] s共t兲 ⫽ [153]


u⫽1

where the two exponents were combined into the


phase ⌽ uv (t, t 0 ), defined as the difference between Next we use the idea that the basis set is orthonormal
the dynamic phases of states 兩v典 and 兩u典, and insert the closure identity (Eq. [112]):
144 EDÉN

Time-Independent Hamiltonian

s共t兲 ⫽ 具u兩␳ˆ 共t兲兩v典具v兩 Q̂兩u典 [154] Because the Hamiltonian has time-independent eigen-
values, the expression for the dynamic phase ⌽ uv (t,
u,v⫽1

0) evaluates to the simple expression


which may be compared with a similar expression,
Eq. [134], that employed an arbitrary basis set. After t


inserting Eq. [118], together with the analytical ex-
pression for the dynamically inhomogeneous propa- ⌽ uv共t, 0兲 ⫽ dt⬘共␻ v ⫺ ␻ u兲 [160]
gator (Eq. [145]), the signal may be expressed as
0

冘 具u兩Û共t, 0兲␳ˆ 共0兲Û共t, 0兲 兩v典具v兩Q̂兩u典


s共t兲 ⫽ †
[155] ⫽ 共␻ v ⫺ ␻ u兲 䡠 t [161]
u,v⫽1 ␻uv

冘 where we defined the frequency ␻ uv as the eigenvalue


⫽ 具u兩 䡠 exp兵⫺i⌽r 共t, 0兲其兩r典具r兩 [156] difference,


u,v,r,s⫽1 Û共t, 0兲

⫻ ␳ˆ 共0兲 䡠 exp兵i⌽s 共t, 0兲其兩s典具s兩 䡠 兩v典具v兩Q̂兩u典 ␻ uv ⫽ ␻ v ⫺ ␻ u [162]

Û共t, 0兲 †
By defining an amplitude as the product of the matrix
elements of the initial density operator and observ-
Rearranging the exponential factors and using Eq. able,
[152] gives
a uv ⫽ 具u兩␳ˆ 共0兲兩v典具v兩Q̂兩u典 [163]

s共t兲 ⫽ 具r兩␳ˆ 共0兲兩s典 䡠 具v兩Q̂兩u典 [157]


u,v,r,s⫽1
Equation [159] casts as (2, 4)

⫻ exp兵i⌽rs 共t, 0兲其 䡠 具u兩r典 䡠 具s兩v典


冘 a exp兵i␻ t其

s共t兲 ⫽ uv uv [164]
Equating the scalar products 具u兩r典 and 具s兩v典 with the u,v⫽1
Kronecker delta functions ␦(u, r) and ␦(s, v), respec-
tively, provides the following expression for the time Thus, the time-domain signal from a time-indepen-
signal dent Hamiltonian is given by a sum of complex ex-
ponentials oscillating at the angular frequency differ-

冘 ences ␻ uv of the Hamiltonian eigenvalues.


s共t兲 ⫽ 具r兩␳ˆ 共0兲兩s典 The time-domain signal is converted into the NMR
u,v,r,s⫽1 spectrum by a Fourier transform
⫻ 具v兩Q̂兩u典exp兵i⌽rs 共t, 0兲其␦共u, r兲␦共s, v兲 [158]

冘 a ␦共␻, ␻ 兲

S共␻兲 ⫽
冘 具u兩␳ˆ 共0兲兩v典具v兩Q̂兩u典exp兵i⌽ 共t, 0兲其
ᏺ uv uv [165]
⫽ uv [159] u,v⫽1

u,v⫽1

Equation [165] shows that the spectrum from a time-


Equation [159] is the key result of this section: it independent Hamiltonian is represented by a set of
corresponds to the general expression for the NMR “spikes” (␦ functions) at the frequency coordinates
time-domain signal from a spin system evolving under ␻ ⫽ ␻ uv . The corresponding amplitudes a uv are
a dynamically inhomogeneous Hamiltonian. Below given as products of the elements of the observable
we examine how it is evaluated in two limiting cases: and initial density operator, both expressed in the
when the Hamiltonian is either time-independent or eigenbasis of the Hamiltonian. An experimental spec-
periodic in time. These correspond to the situation of trum always has “broad” peaks due to decay mecha-
static and rotating samples, respectively, and they are nisms of the time signal (“relaxation”). Assuming an
the scenarios considered in the numerical simulations exponential decay of the time-domain signal, the cor-
in subsequent articles. responding spectrum that results after Fourier trans-
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 145

density operator proportional to Îx , which in turn may


be expressed in terms of Î⫹ and Î⫺ using Eq. [75],

␳ˆ 共0兲 ⫽ Î x [166]
1 ⫹
⫽ 共Î ⫹ Î⫺ 兲 [167]
2

or, formulated in the Dirac formalism,

1
␳ˆ 共0兲 ⫽ 共兩␣典具␤兩 ⫹ 兩␤典具␣兩兲 [168]
2

We assume for simplicity that the spins are only


affected by the isotropic chemical shift interaction,
corresponding to the Hamiltonian

Figure 11 A typical spectrum from (a) time-independent Ĥ CS,iso ⫽ ␻isoÎz [169]

冉 冊
and (b) time-periodic Hamiltonians, corresponding to the
case of one single molecular orientation (“single crystal”) in 1 1 0
⫽ ␻ iso 0 ⫺1 [170]
static and rotating samples, respectively. Only one transition 2
of the spin system is shown, involving the Hamiltonian
eigenstates {兩u典, 兩v典}. In the static case, the pair of eigen- where we inserted the explicit matrix representation
states generates a single spectral peak at the frequency ␻ uv of Îz (Eq. [77]). Because the Hamiltonian is diagonal
(given by the difference between the eigenvalues of the in the Zeeman basis, its eigenvalues are directly ex-
states: ␻ uv ⫽ ␻ v ⫺ ␻ u ) and with the amplitude a uv (Eq. tracted as ␻␣ ⫽ ␻iso /2 and ␻␤ ⫽ ⫺␻iso /2. Hence, the
[163]). In the rotating case, the spectrum corresponds to a
eigenequations are
side-band manifold, centered at ␻ uv (0)
⫽ ␻ v(0) ⫺ ␻ (0)
u , and
with the expressions for the frequencies ␻ uv(k)
and amplitudes
(k)
a uv of the side bands given by Eqs. [212] and [211], 1
Ĥ CS,iso兩␣典 ⫽ ␻iso兩␣典 [171]
respectively. Two neighboring side bands are separated by 2
the rotational frequency ␻ r . In this case, we assumed that
␻ uv ⫽ ␻ uv(0)
, but this does not generally hold, as explained 1
in the text.
Ĥ CS,iso兩␤典 ⫽ ⫺ ␻iso兩␤典 [172]
2

The eigenvalues are time independent, and the


formation contains peaks with Lorentzian shapes in- expressions for the corresponding dynamic phases
⌽ ␣ (t, 0) and ⌽ ␤ (t, 0) are easily calculated from Eq.
stead of ␦ functions. Line broadening may also be
[143]:
introduced in simulated spectra, as will be explained
in the next article. Figure 11(a) shows a typical spec-
t


trum obtained from one of the terms in Eq. [165].
1
Example: Isotropic Chemical Shift Evolution. To ⌽ ␣共t, 0兲 ⫽ ⫺⌽␤ 共t, 0兲 ⫽ ␻iso dt⬘ [173]
2
illustrate spin evolution under a time-independent
0
Hamiltonian, we explicitly derive the form of the
NMR time-domain signal generated from the isotro- 1
pic chemical shift interaction using the general for- ⫽ ␻ t [174]
2 iso
malism presented earlier. We will show that the result
obtained here is consistent with the expressions for Applying Eq. [152] to form the dynamic phase dif-
the time signal derived above (Eq. [164]). ferences ⌽ ␣␤ (t, 0) and ⌽ ␤␣ (t, 0) gives
Assume an ensemble of isolated spins-21 prepared
in a state of transverse magnetization, generated by a ⌽ ␣␤共t, 0兲 ⫽ ⌽ ␤共t, 0兲 ⫺ ⌽ ␣共t, 0兲 [175]
(␲/ 2) y pulse at the start of signal acquisition (t 0 ⫽
0), as discussed earlier. This corresponds to an initial ⫽ ⫺␻isot [176]
146 EDÉN

and respectively. Moreover, the sum over j collapses into


a single term
⌽ ␤␣共t, 0兲 ⫽ ⫺⌽␣␤ 共t, 0兲 ⫽ ␻isot [177]
1
s共t兲 ⫽ exp兵i␻isot其
respectively. 2

冉 冊
Following reasoning identical to the previous sec-
tion, we obtain an equivalent expression to Eq. [151]: ⫻ 具␣兩 䡠 兩␤典具␤兩 䡠 兩␣典 ⫹ 具␤兩 䡠 兩␤典具␤兩 䡠 兩␤典 [184]

⫽0 ⫽1
1
␳ˆ 共t兲 ⫽ 共exp兵i⌽␣␤ 共t, 0兲其兩␣典具␤兩 [178] 1
2
⫽ exp兵i␻isot其 [185]
2
⫹exp兵i⌽␤␣ 共t, 0兲其兩␤典具␣兩兲
1 Note that all contributions from the ⫹1QC operator
⫽ 共exp兵⫺i␻isot其兩␣典具␤兩 [179] 兩␣典具␤兩 vanished; this is the mathematical consequence
2
of the fact that quadrature detection implies detection
⫹ exp兵i␻isot其兩␤典具␣兩兲 of ⫺1QC. Thus, in the present case, the signal only
contains one frequency: the isotropic chemical shift
According to Eq. [179], the density operator at the ␻iso. Subsequent Fourier transformation produces an
time point t is given by two terms, each being a NMR spectrum corresponding to a single peak at the
product of an operator and a time-dependent expo- isotropic chemical shift frequency, as indicated in Fig.
nential function involving a dynamic phase. The 11(a). Note that Eq. [185] conforms to Eq. [164] with
physical meaning of these terms is that the ⫹1Q the following expressions for the frequency and am-
coherence (represented by 兩␣典具␤兩) oscillates at the plitude: ␻ uv ⫽ ␻ iso and a uv ⫽ 1/ 2.
negative value of the isotropic chemical shift ⫺␻iso, This example represented probably the simplest
whereas the ⫺1Q coherence (represented by 兩␤典具␣兩) possible case. However, the calculations outlined in
oscillates at ⫹␻iso. the next section extend the present case by allowing
The time-domain signal originating from the time the incorporation of anisotropic chemical shift contri-
evolution of this spin ensemble is obtained by insert- butions and heteronuclear dipolar interactions in mul-
ing the expression for the density operator together tispin systems. In all these cases, the spin Hamiltonian
with the observable Î⫹ ⫽ 兩␣典具␤兩 into Eq. [133]: is dynamically inhomogeneous and, furthermore, di-
agonal in the Zeeman product basis. See Ref. (36) for
a similar calculation that additionally includes the
s共t兲 ⫽
1
2
冘 具 j兩␳ˆ Q̂兩 j典
j⫽␣,␤
[180] anisotropic chemical shift interaction.


1
2
冘 具 j兩 䡠 冉exp兵⫺i␻
j⫽␣,␤
t其兩␣典具␤兩 䡠 兩␣典兩␣典具␤兩
iso
Time-Periodic Self-Commuting
Hamiltonian

Here we carry out analogous calculations for a rotat-

⫹ exp兵i␻isot其兩␤典具␣兩 䡠 兩␣典具␤兩 䡠 兩 j典

冊 [181]
ing sample as was done for the static case in the
previous section. The key steps of the procedure are
the same: (i) obtain the Hamiltonian and its eigenval-
ues; (ii) calculate the expression for the time-depen-
Again, the orthonormality condition of the basis states dent dynamic phase ⌽ uv (t, 0) originating from each
proves useful: the first and second product within the pair of eigenvalues; (iii) insert these results into the
parentheses may be simplified as expression for the time-domain signal, Eq. [159].

Hamiltonian. The time-dependent dynamically inho-


兩␣典具␤兩 䡠 兩␣典具␤兩 ⫽ 0 [182]
mogeneous Hamiltonian is given by a sum over all
NMR interactions:
and

兩␤典具␣兩 䡠 兩␣典具␤兩 ⫽ 兩␤典具␤兩 [183]


Ĥ共t兲 ⫽ 冘 Ĥ 共t兲 ⫽ 冘 ␻ 共t兲T̂



⌳ ⌳ [186]
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 147

where T̂⌳ represents the spin part of the interaction


冘␻
2
共m兲
and the various spin operators mutually commute: ␻ u共t兲 ⫽ u exp兵im␻r t其 [192]
[T̂⌳ j, T̂⌳ k] ⫽ 0. If two operators commute, they share m⫽⫺2
the same eigenbasis and may both be brought into
diagonal form in that basis (although the eigenvalues An explicit form of the Fourier components ␻ (m)
u may
of the operators are generally different) (29). From be obtained by inserting Eq. [104] into Eq. [190] as
this it follows that all operators in the sum of Eq. follows:
[186] are diagonal in the Hamiltonian eigenbasis.
Inserting the closure relation (Eq. [112]) both to the
冘冘␻
2
left and to the right of Eq. [186] gives 共m兲
␻ u共t兲 ⫽ ⌳ 具u兩T̂ ⌳ 兩u典exp兵im␻r t其 [193]
⌳ m⫽⫺2

冘 冘 ␻ 共t兲兩v典具v兩 䡠 T̂
ᏺ ␻ 共m兲
u

Ĥ共t兲 ⫽ ⌳ ⌳ 䡠 兩v典具v兩 [187]


v⫽1 ⌳ and equating the Fourier coefficients with those of Eq.
1̂ 1̂
[192]. This is possible since a Fourier series is unique,

冘 冘 ␻ 共t兲具v兩T̂ 兩v典 䡠 兩v典具v兩


ᏺ meaning that the Fourier coefficients of the two ex-
⫽ ⌳ ⌳ [188] pansions must be equal, providing the following re-
v⫽1 ⌳ lationship between the Fourier components of the
eigenvalues and those of the spatial tensors:
Note that 具v兩T̂⌳ 兩v典 represents the matrix element
(T̂⌳ ) vv . ␻ u共m兲 ⫽ 冘␻

共m兲
⌳ 具u兩T̂ ⌳ 兩u典 [194]

Hamiltonian Eigenvalues. Next we derive the ex-


plicit expressions for the Hamiltonian eigenvalues. Thus, the eigenvalue Fourier component ␻ (m)
u is given
We may project out the uth Hamiltonian eigenvalue by the sum over the products of the mth Fourier
by multiplying Eq. [188] to the left with the “bra” 具u兩 component from each spatial tensor and the matrix
and to the right with the “ket” 兩u典, according to element of the corresponding spin tensor operator,
expressed in the eigenbasis of the Hamiltonian.
␻ u共t兲 ⫽ 具u兩Ĥ共t兲兩u典
Dynamic Phase. Next we deal with finding an ex-

⫽ 具u兩 䡠 再冘 冘

v⫽1 ⌳
␻ ⌳共t兲具v兩T̂ ⌳ 兩v典 䡠 兩v典具v兩 䡠 兩u典 冎 pression for the dynamic phase ⌽ uv (t, t 0 ). To this
end, we separate the Hamiltonian eigenvalue in Eq.
[192] into time-independent and time-dependent parts
as follows:
[189]

After rearranging the terms and using the orthonor-


␻ u共t兲 ⫽ ␻ u共0兲 ⫹ 冘␻
m⫽0
共m兲
u exp兵im␻r t其 [195]
mality conditions of the Hamiltonian eigenstates, all
terms for which v ⫽ u vanish, and we obtain
Under MAS conditions, ␻ (0) u only contains contribu-

冘 ␻ 共t兲具u兩T̂ 兩u典
tions from the isotropic parts of the spin interactions
␻ u共t兲 ⫽ ⌳ ⌳ [190] (represented by zeroth rank tensors) whereas the com-
⌳ ponents ␻ (m⫽0)
u contain only anisotropic parts (repre-
sented by components of second rank tensors). The
As discussed earlier, the anisotropy frequency reason that ␻ (0)
u is purely isotropic is that all potential
␻ ⌳ (t) of the spatial parts of the Hamiltonian are time contributions from second rank tensors are zero: this
periodic in a sample rotating at the constant frequency follows from Eq. [194] and the fact that d 200 (␪ m ) ⫽ 0
␻ r , Eq. [190] then implies that the Hamiltonian eig- in the expression for ␻⌳ (0)
(Eq. [105]). However, for
envalues are also periodic with the same period ␶ r : experiments conducted under off-MAS conditions,
the Euler angle ␤ RL ⫽ ␪ m and anisotropic contribu-
tions will appear in the frequency ␻ (0)
u . In the follow-
␻ u共t ⫹ ␶ r兲 ⫽ ␻ u共t兲 [191] ing, we ignore such cases.
As discussed below, the separation between isotro-
They may accordingly be expanded in a Fourier series: pic and anisotropic contributions to the eigenvalues
148 EDÉN

and, subsequently, the dynamic phase, will have im-


plications for the nature of the NMR time-domain
signals and frequency-domain spectra obtained from
rotating solids.
By combining Eqs. [143] and [195], we obtain the
following expression for the dynamic phase ⌽ u (t, t 0 ):

⌽ u共t, t 0兲 ⫽ 冕 dt⬘␻ u共t⬘兲 [196]

t0

t t

冕 冕
Figure 12 Diagrams illustrating each of the five Fourier
⫽ dt⬘␻ u共0兲 ⫹ 冘
m⫽0
␻ u共m兲 dt⬘exp兵im␻r t⬘其 components ␻ ⌳ (m)
(Eq. [105]) of a spatial tensor in a rotating
solid at a given time point t. Each circle corresponds to the
t0 t0
plane of complex numbers, and the vectors represent the
Fourier components. The component ␻ ⌳ (m)
rotates at the
␻共0兲
u 共t⫺t0兲
⌽⬘u 共t, t 0 兲 velocity m␻ r . Note that the magnitudes (i.e., the lengths of
the vectors) are equal for two components with opposite
[197] signs of m, but they rotate in opposite directions. At t ⫽ 0,
all vectors are aligned along the real axis; and at every
As a consequence of the separation of the Hamilto- integer multiple of the rotational period, they return to the
nian eigenvalues into time-independent and time-de- initial position. The angle ⌽ ⌳ (m)
represents the contribution
pendent parts, the dynamic phase ⌽ u (t, t 0 ) involves to the dynamic phase accumulated by the component ␻ ⌳ (m)
.
the constant frequency ␻ (0) The dynamic phase of a Hamiltonian eigenstate ⌽ u (Eq. (m)
u and a time-dependent part
⌽⬘u (t, t 0 ): [198]) is the sum of contributions from all interactions. The
m ⫽ 0 component ␻⌳ (0)
is independent of time and stays
aligned along the real axis at all times. Its dynamic phase
⌽ u共t, t 0兲 ⫽ ␻ u共0兲共t ⫺ t 0兲 ⫹ ⌽⬘u共t, t 0兲 [198] may be visualized as the area beneath the line y ⫽ ␻ ⌳ (0)
over
the interval 0 ⱕ t⬘ ⱕ t.
The integration of the exponential functions in ⌽⬘u (t,
t 0 ) are easily carried out,

t
responding dynamic phases for a single interaction ⌳


is given in Fig. 12; the dynamic phase ⌽ u (t, t 0 ) of the
Hamiltonian eigenstate 兩u典 is a linear combination of
dt⬘exp兵im␻r t⬘其
the dynamic phases of the individual interactions
t0 ⌽ ⌳ (t, t 0 ), with the corresponding Fourier compo-
nents ␻ (m)
u and ␻ ⌳
(m)
related through Eq. [194].
⫽ 共im␻ r兲 ⫺1兵exp兵im␻r t其 ⫺ exp兵im␻r t0 其其 [199] Next we demonstrate that ⌽⬘u (t, t 0 ) is periodic:

and the following expression for the phase ⌽⬘u (t, t 0 ) is ⌽⬘u共t ⫹ ␶ r, t 0兲 ⫽ ⌽⬘u共t, t 0兲 [201]
obtained:

冘m
This may be verified from Eq. [200] by similar argu-
⌽⬘u共t, t 0兲 ⫽ 共i␻ r兲 ⫺1 ⫺1
␻ u共m兲 ments that were used to demonstrate that the interac-
m⫽0 tion frequency ␻⌳
(t)
is periodic. From Eq. [200] we get
⫻ (exp兵im␻r t其 ⫺ exp兵im␻r t0 其) [200]
⌽⬘u共t ⫹ ␶ r, t 0兲 ⫽ 共i␻ r兲 ⫺1 冘m
m⫽0
⫺1
␻ u共m兲
Note that ⌽⬘u (t, t 0 ) only contains contributions from
the anisotropic parts of the spin interactions, because ⫻ 共exp兵im␻r 共t ⫹ ␶r 兲其 ⫺ exp兵im␻r t0 其兲 [202]
Eq. [200] solely comprises components ␻ (m) u with
m ⫽ 0: these may only originate from the anisotropic Because the product of the rotational period and the
interaction parts expressed by second rank tensors. An spinning frequency is given by ␻ r ␶ r ⫽ 2␲, we find
interpretation of the Fourier components and the cor- exp{im␻ r (t ⫹ ␶ r )} ⫽ exp{im␻ r t}exp{2im␲} ⫽
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 149

exp{im␻ r t}; Eq. [202] reduces to Eq. [200], imply- harmonics of the rotational frequency ␻ r . This may be
ing that ⌽⬘u (t ⫹ ␶ r , t 0 ) ⫽ ⌽⬘u (t, t 0 ). realized by Fourier expanding the function
Finally, we seek an explicit expression for the exp{i⌽⬘uv (t, 0)}: as expressed by Eq. [27], the com-
dynamic phase difference ⌽ uv (t, t 0 ). Inserting Eq. plex exponentials of periodic functions (such as the
[198] (which separates the isotropic and anisotropic dynamic phase) may be written as a Fourier series,
contributions of ⌽uv(t, t 0 ) into Eq. [152] gives comprising an infinite number of coefficients:

⌽ uv共t, t 0兲 ⫽ 兵␻ v共0兲共t ⫺ t 0兲 ⫹ ⌽⬘v共t, t 0兲其


冘c

共k兲
exp兵i⌽⬘u v 共t, 0兲其 ⫽ uv exp兵ik␻r t其 [208]
⫺兵␻u共0兲 共t ⫺ t0 兲 ⫹ ⌽⬘u 共t, t0 兲其 [203] k⫽⫺⬁

共0兲
⫽ ␻ uv 共t ⫺ t 0兲 ⫹ ⌽⬘uv共t, t 0兲 [204] (k)
The Fourier coefficient c uv is a function of the Ham-
iltonian eigenvalue difference ␻ (m) ⫺ ␻ (m) and hence
Analogous to the frequency ␻ uv in Eq. [162], which is v u
also depends on the various spatial tensor compo-
relevant for the static solid, we define the frequency
nents; this is realized from the relationships between
␻ uv
(0)
as the difference between the m ⫽ 0 Fourier
the dynamic phase, the Hamiltonian eigenvalues, and
components of the Hamiltonian eigenvalues:
the interaction parameters, as follows by combining
(k)
共0兲
Eqs. [105], [194], and [206]. The coefficients c uv
␻ uv ⫽ ␻ v共0兲 ⫺ ␻ u共0兲 [205] fulfill a normalization condition ¥ k 兩c uv 兩 ⫽ 1 and
(k)

may be calculated by Fourier transformation of the


Moreover, an explicit expression for the phase differ- function exp{i⌽⬘uv (t, 0)}. We will discuss this fur-
ence ⌽⬘uv (t, t 0 ) ⫽ ⌽⬘v (t, t 0 ) ⫺ ⌽⬘u (t, t 0 ) is obtained ther in following articles.
by using Eq. [200]: Finally, by inserting Eqs. [207] and [208] into the

冘m
general expression for the signal, Eq. [159], we obtain
⌽⬘uv共t, t 0兲 ⫽ 共i␻ r兲 ⫺1 ⫺1
兵␻ v共m兲 ⫺ ␻ u共m兲其

冘 具u兩␳ˆ 共0兲兩v典具v兩Q̂兩u典exp兵i␻
m⫽0 ᏺ
共0兲
⫻ 共exp兵im␻r t其 ⫺ exp兵im␻r t0 其兲 [206] s共t兲 ⫽ uv t其 [209]
u,v⫽1

NMR Responses in Time and Frequency Domains.


冘c

共k兲
Equipped with the expressions derived in the previous ⫻ uv exp兵ik␻r t其
subsections, we are prepared for the final step: eval- k⫽⫺⬁
uating the time signal, Eq. [159], in the case of a

冘 冘a
rotating solid with a self-commuting Hamiltonian. As ᏺ ⬁
共k兲 共k兲
before, the start of the NMR signal acquisition defines ⫽ uvexp兵i␻uv t其 [210]
the time origin, t 0 ⫽ 0. u,v⫽1 k⫽⫺⬁

Exponentiating Eq. [204] provides the following


expression for exp{i⌽ uv (t, 0)}: where the amplitudes are given by

共k兲 共k兲 共k兲


共0兲
exp兵i⌽uv 共t, 0兲其 ⫽ exp兵i␻uv t其exp兵i⌽⬘u v 共t, 0兲其 [207] a uv ⫽ 具u兩␳ˆ 共0兲兩v典具v兩Q̂兩u典cuv ⫽ auv cuv [211]

It corresponds to a product of the exponentiated eig- and the frequencies by


envalue difference ␻ uv(0)
and the exponential of the
共k兲 共0兲
periodic anisotropic phase ⌽⬘uv (t, t 0 ). The former is ␻ uv ⫽ ␻ uv ⫹ k␻ r [212]
analogous to the factor exp{i␻ uv t} from the static
case; both functions comprise a single frequency. After Fourier transformation of the time signal, the
Note, however, that the frequencies ␻ uv and ␻ uv (0)
are spectrum takes the following form:
in general not equal. The frequency ␻ uv contains both

冘 冘a
isotropic and anisotropic contributions from the spin ᏺ ⬁
共k兲 共k兲
interactions, but in the MAS case we arranged that S共␻兲 ⫽ uv␦共␻, ␻ uv 兲 [213]
␻ uv
(0)
only comprises isotropic interactions. u,v⫽1 k⫽⫺⬁

The factor exp{i⌽⬘uv (t, 0)} has no counterpart in


the case of a static sample; in general, this function Consequently, the spectrum is the sum of contribu-
comprises an infinite number of frequencies, all being tions from each pair of Hamiltonian eigenstates 兩u典
150 EDÉN

and 兩v典, each producing a set of spinning side bands tween the various noncommuting terms in the total
due to the periodic sample rotation. The side-band Hamiltonian (as discussed on page 141).
manifold is centered at a “fundamental frequency” We summarize by comparing the expressions for
␻ uv
(0)
and associated with a frequency separation ␻ r the NMR time-domain signal and spectrum obtained
between neighboring peaks. This is illustrated in Fig. from a time-independent Hamiltonian (Eqs. [164] and
11(b). [165]) with those from a periodic Hamiltonian (Eqs.
The origin of the side bands in the rotating case is [210] and [213]): in the former case, each pair of
the periodic exponential function exp{i⌽⬘uv (t, 0)}. states {兩u典, 兩v典} produces a single spectral peak [Fig.
Through Eq. [208] it was shown to correspond to a 11(a)], whereas in the latter case each state pair is
sum of functions exp{ik␻ r t}, each oscillating at a associated with a manifold of spectral peaks [Fig.
(k)
harmonic k␻ r of the spinning frequency. Each of 11(b)]. Each amplitude a uv is the product of a uv
these contributes one spinning side band to the man- (from the time-independent case) and the Fourier
(k)
ifold. However, although the number of side bands is component c uv , as expressed by Eq. [211]. The sum
(k)
in theory infinite, in practice their amplitudes, a uv ⫽ over all side-band amplitudes a uv (k)
within a manifold
a uv c uv , tend to zero for large side-band orders 兩k兩.
(k)
equals a uv .
(k)
This is because the Fourier components c uv get
smaller as the side-band index 兩k兩 increases: the num-
SUMMARY
ber of side bands of significant intensity roughly
depends on the magnitude of the interaction relative to
We have discussed a framework for calculating the
the spinning frequency, that is, on the ratio 兩␻ ⌳ /␻ r 兩
NMR time-domain signal and frequency-domain
(27). The larger the ratio, the larger the number of
spectrum generated from a single molecular orienta-
side bands in the spectrum. This is seen, for example,
tion in solid-state NMR. The key steps are the fol-
from the experimental spectra in Fig. 1(b– d), which
lowing:
were recorded at different spinning frequencies. In the
case of several interactions, the ratio involving the 1. Construct the Hamiltonian for the spin system.
sum over all interaction frequencies, ␻⌺ ⫽ ¥⌳ ␻⌳, This is given as a product of a spatial tensor and
roughly dictates the number of side bands. a spin operator, each represented by a compo-
Note that spin evolution under a self-commuting nent of an irreducible spherical tensor. We
Hamiltonian of the form of Eq. [186] corresponds to showed how the spatial tensor may be trans-
an NMR spectrum being a sum of a set of “subspec- formed between different reference frames and
tra,” each generated independently from each interac- how the matrix representation for the spin op-
tion. For example, consider an isolated 13CO1H seg- erator may be calculated.
ment in an organic molecule under MAS. The 13C 2. The density operator carries the information
spectrum corresponds to a side band manifold, with about the state of an ensemble of nuclear spin
each side band amplitude arising from the combined systems. The Hamiltonian will cause the den-
effects of the heteronuclear dipolar coupling and the sity operator to change throughout the NMR
13
C chemical shift anisotropy. In a following article experiment. The fundamental problem in spin
we demonstrate how such a spectrum may be calcu- dynamics calculations is solving the Schröd-
lated numerically. inger equation, which dictates how the spin
The derivations in this section assumed dynami- density operator, and hence the nuclear spins,
cally inhomogeneous Hamiltonians under sample evolves in time; this results in an operator called
spinning conditions, but the form of the spectrum the propagator. In practice, the propagator usu-
given in Eq. [213] holds in the general case of spin ally has to be estimated numerically. It may
systems evolving under any time-periodic Hamilto- then be used to calculate the spin density oper-
nian, as discussed in Refs. (18, 20, 39 – 41). The ator at any time point during the NMR experi-
appearance of a side-band manifold is a consequence ment.
of spin evolution under a time-periodic Hamiltonian, 3. The time-domain signal s(t) at a given time
regardless of whether this is homogeneous or inho- point corresponds to the expectation value of an
mogeneous. The major difference between these cases observable operator. It is calculated as the trace
is, however, that for dynamically homogeneous Ham- of the product of the observable operator and
iltonians, the total NMR spectrum generated from the density operator at the given time point.
several interactions is not expressible as a superposi-
tion of the individual spectra obtained from each By repeating the above steps for a series of time
interaction. This is attributable to interference be- points, one obtains the time-domain signal “traject-
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 151

ory” s(t). The corresponding NMR spectrum is cal-



2

culated by Fourier transformation of s(t). We derived ⫽ 共⫺1兲⫺m⬘ 关A2m⬘ 兴M D⫺m⬘m
2
共⍀MR 兲*dm0
2
共␪m 兲
the formal equations describing these steps and ex- m⬘⫽⫺2
amined in particular detail the NMR response ob-
[A.4]
tained if the spin Hamiltonian is dynamically inho-
mogeneous (i.e., self-commuting at all times during
where the two factors (⫺1) ⫺m⫺m⬘ and (⫺1) ⫺m were
the NMR signal acquisition). This allowed the spin
combined into
dynamics to be solved exactly. Finally, the form of the
spectrum obtained from a single molecular orientation
in a static and rotating sample were examined. 共⫺1兲⫺m 䡠 共⫺1兲⫺m⫺m⬘ ⫽ 共⫺1兲⫺2m⫺m⬘ ⫽ 共⫺1兲⫺m⬘
The next article will outline how this formalism is [A.5]
converted into computer code in order to numerically
calculate NMR signals. and the last equality holds for integral values of m.
After a change of index n ⫽ ⫺m⬘, the expression for
␻⌳(⫺m)
is

ACKNOWLEDGMENTS
冘 共⫺1兲 关A
2
共⫺m兲 ⌳
␻ ⌳ ⫽ n
2⫺n
M
兴 Dnm
2
共⍀MR 兲*dm0
2
共␪m 兲 [A.6]
I would like to thank L. Frydman, C.V. Grant, A. n⫽⫺2
Sebald, and S. Vega for helpful comments on the
manuscript and M.H. Levitt for many discussions and From Eq. [51] it follows that (⫺1) n [A ⌳ M

2⫺n ]
for providing Mathematica routines used for generat- ⌳ M
([A 2n ] )*. Because the reduced Wigner functions are
ing some of the figures. A postdoctoral fellowship 2
real numbers, that is, d m0 (␪ m ) ⫽ d m0
2
(␪ m )*, Eq. [106]
from The Swedish Foundation for International Co- follows by using standard properties of complex num-
operation in Research and Higher Education (STINT) bers:
is appreciated.

冘 共⫺1兲 关A
2

␻ ⌳共⫺m兲 ⫽ n ⌳
2⫺n
M
兴 䡠 Dnm
2
共⍀MR 兲* 䡠 dm0
2
共␪m 兲*
n⫽⫺2
APPENDIX A ⌳ M
共关A 2n 兴 兲*

冉冘 冊
We prove (i) the symmetry of the Fourier components [A.7]
*
␻⌳(m)
upon sign reversal of m (Eq. [106]), (ii) the 2
⌳ M 2
expression for the spatial tensor component ␻ ⌳ (t) ⫽ ⫽ 关A 2n 兴 D nm共⍀ MR兲d m0
2
共␪ m兲
[A.8]
[A ⌳
20 ]
L
(Eq. [107]), and (iii) that the latter is real n⫽⫺2

valued. These properties follow from the relationship


of the irreducible spherical tensor components (Eq. ⫽ ␻ ⌳共m兲* [A.9]
[51]) and the following symmetries of the Wigner
functions (28): Once Eq. [106] is established, it is easy to show that
␻ ⌳ (t) is real: we rearrange the sum in Eq. [104] as
l
d m⬘m共␤兲 ⫽ 共⫺1兲m⬘⫺m d ⫺m⬘⫺m
l
共␤兲 [A.1] follows:
l
D m⬘m共⍀兲 ⫽ 共⫺1兲m⫺m⬘ D⫺m⬘⫺m
l
共⍀兲* [A.2]
冘␻
2
共m兲
␻ ⌳共t兲 ⫽ ⌳ exp兵im␻r t其 [A.10]
m⫽⫺2
These equations may be checked explicitly using the
Wigner functions given in Table 1. Substituting them
冘␻
2
into the the expression for ␻ ⌳
(⫺m)
(Eq. [105]) gives
⫽ ␻ ⌳共0兲 ⫹ 共m兲
⌳ exp兵im␻r t其
m⫽1


2
共⫺m兲 ⌳ M
␻ ⌳ ⫽ 关A 2m⬘ 兴 䡠 2
D m⬘⫺m 共⍀ MR兲 䡠 d⫺m0
2
共␪m 兲 ⫹␻⌳(⫺m) exp兵⫺im␻r t其 [A.11]
m⬘⫽⫺2
2 2
共⫺1兲 ⫺m⫺m⬘ D ⫺m⬘m 共⍀ MR 兲* 共⫺1兲 ⫺m d m0 共␪ m 兲
Using Eq. [106], as well as exp{⫺im␻ r t} ⫽
[A.3] exp{im␻ r t}*, we obtain
152 EDÉN

We focus on the second-order term. It may be for-



2

␻ ⌳共t兲 ⫽ ␻ ⌳共0兲 ⫹ ␻ ⌳共m兲exp兵im␻r t其 mally expressed as a sum over ᏺ2 products of pro-


m⫽1 jection operators according to

冉 冊
⫹ ␻⌳共m兲* exp兵im␻r t其* [A.12] 2

冘 冘
ᏺ ᏺ
1 1
⫺i ⌽u 兩u典具u兩 ⫽ ⌽ ⌽ 兩u典具u兩 䡠 兩v典具v兩
Note that for each value of m, we have a sum of a 2! u⫽1
2! u,v⫽1 u v
complex number ␻⌳ (m)
exp{im␻r t} and its complex [B.3]
(m) (m)
conjugate (␻⌳ exp{im␻r t})*; this equals 2 Re(␻⌳
exp{im␻r t}). Because it directly follows from Eq.
This appears at first sight to be a very complex ex-
[106] that ␻⌳(0)
is real, Eq. [225] corresponds to a
pression. However, by using the properties of “bras”
sum of three real numbers:
and “kets” and considering that the Hamiltonian ei-
genstates are orthonormal, each operator product on

冘 Re共␻
2 the right-hand side may be evaluated as
␻ ⌳共t兲 ⫽ ␻ ⌳共0兲 ⫹ 2 共m兲
⌳ exp兵im␻r t其兲 [A.13]
m⫽1 兩u典具u兩 䡠 兩v典具v兩 ⫽ 兩u典 䡠 具u兩v典 䡠 具v兩 ⫽ 兩u典具v兩 䡠 ␦共u, v兲
␦共u, v兲
Finally, Eq. [107] is established by evaluating the real
part of the product ␻ ⌳(m)
exp{im␻ r t}, resulting in [B.4]

The result is that (兩u典具u兩) 2 ⫽ 兩u典具u兩, while all prod-

冘 Re兵␻
2
ucts between different operators vanish. This may be
␻ ⌳共t兲 ⫽ ␻ ⌳共0兲 ⫹ 2 共m兲
⌳ 其cos兵m␻r t其 extrapolated to obtain the following expression for the
m⫽1
nth-order term in Eq. [B.2]:

冉 冊
⫺ Im兵␻⌳共m兲 其sin兵m␻r t其 [A.14]
N

冘 冘
ᏺ ᏺ
1 1
⫺i ⌽u 兩u典具u兩 ⫽ 共⫺i⌽u 兲N 兩u典具u兩
N! u⫽1
N! u⫽1
APPENDIX B
[B.5]
Following Ref. (36), we show that in the Hamiltonian
eigenbasis, the propagator may be expressed as a sum It follows that all terms of index u in Eq. [B.2] are
of products of projection operators and exponentials proportional to the projection operator 兩u典具u兩; this
of dynamic phases (Eq. [145]). Here we use the short- makes the following factorization possible:
hand notation ⌽ u ⬅ ⌽ u (t, t 0 ) and insert the expres-

冘 再1̂ ⫹ 共⫺i⌽ 兲 ⫹ 2!1 共⫺i⌽ 兲


sion for the dynamic phase (Eq. [143]) into Eq. [142]: ᏺ

Û共t, t0 兲 ⫽

再 冎
2
u u

冘 ⌽ 兩u典具u兩
ᏺ u⫽1


Û共t, t0 兲 ⫽ exp ⫺i u [B.1]
1
u⫽1
⫹ · · ·⫹ 共⫺i⌽u 兲N ⫹ · · · 兩u典具u兩 [B.6]
N!
According to Eq. [13], the propagator may be ex-
pressed as the following Taylor series: Inspection of each term with index u in this expres-

冉 冊
sion within braces reveals that it corresponds to the

冘 ⌽ 兩u典具u兩
ᏺ Taylor expansion of the exponential function of the
Û共t, t0 兲 ⫽ 1̂ ⫹ ⫺i u number (⫺i⌽ u ) (see Eq. [12]). Hence, we may re-
u⫽1 place the expression within braces by exp{⫺i⌽ u },

冉 冘 冊
2 whereupon Eq. [B.6] casts as

1
⫹ ⫺i ⌽u 兩u典具u兩 ⫹···⫹
冘 exp兵⫺i⌽ 其兩u典具u兩
2! u⫽1

Û共t, t0 兲 ⫽

冉 冘 冊
u [B.7]
N u⫽1

1
⫺i ⌽u 兩u典具u兩 ⫹··· [B.2]
N! u⫽1 which is the desired result, Eq. [145].
COMPUTER SIMULATIONS IN SOLID-STATE NMR. I 153

REFERENCES of n-alkane molecules enclathrated in urea channels. J


Phys Chem 1996; 100:10854 –10860.
18. Edén M, Lee YK, Levitt MH. Efficient simulation of
1. Haeberlen U. High resolution NMR in solids. Selective
periodic problems in NMR. Application to decoupling
averaging. New York: Academic; 1976.
and rotational resonance. J Magn Reson A 1996; 120:
2. Munowitz M. Coherence and NMR. New York: Wiley;
56 –71.
1988.
19. Dumont RS, Jain S, Bain A. Simulation of many-spin
3. Ernst RR, Bodenhausen G, Wokaun A. Principles of
system dynamics via sparse matrix methodolgy.
nuclear magnetic resonance in one and two dimensions.
J Chem Phys 1997; 106:1–9.
Oxford, UK: Clarendon; 1987.
20. Filip C, Filip X, Demco DE, Hafner S. Spin dynamics
4. Schmidt-Rohr K, Spiess HW. Multidimensional sol-
under magic angle spinning by Floquet theory. Mol
id-state NMR and polymers. New York: Academic;
Phys 1997; 92:757–771.
1994.
21. Studer W. SMART, A general purpose pulse exper-
5. Mehring M, Weberuß VA. Object-oriented magnetic
iment simulation program using numerical density
resonance. Classes and objects, calculations and com-
matrix calculations. J Magn Reson 1988; 77:424 –
putations. London: Academic; 2001.
438.
6. Levitt MH. Spin dynamics. Basics of nuclear magnetic
22. Skibsted J, Nielsen NC, Bildsøe H, Jakobsen HJ. Sat-
resonance. Chichester, UK: Wiley; 2001. ellite transitions in MAS NMR spectra of quadrupolar
7. Andrew ER, Bradbury A, Eades RG. Removal of di- nuclei. J Magn Reson 1991; 95:88 –117.
polar broadening of nuclear magnetic resonance spectra 23. Stickney de Bouregas F, Waugh JS. ANTIOPE, A
of solids by specimen rotation. Nature 1959; 183:1802– program for computer experiments on spin dynamics. J
1803. Magn Reson 1992; 96:280 –289.
8. Lowe IJ. Free induction decays of rotating solids. Phys 24. Smith SA, Levante TO, Meier BH, Ernst RR. Computer
Rev Lett 1959; 2:285–287. simulations in magnetic resonance. An object-oriented
9. Dusold S, Sebald A. Magnitudes and orientations of programming approach. J Magn Reson A 1994; 106:
NMR interactions in isolated three-spin systems ABX. 75–105.
Mol Phys 1998; 95:1237–1245. 25. GAMMA program. Florida State University, Tallahas-
10. Levitt MH, Edén M. Numerical simulation of periodic see, FL. http://gamma.magnet.fsu.edu.
NMR problems: Fast calculation of carousel averages. 26. SIMPSON program. University of Aarhus, Denmark.
Mol Phys 1998; 95:879 – 890. http://nmr.imsb.au.dk.
11. Charpentier T, Fermon C, Virlet J. Efficient time prop- 27. Maricq MM, Waugh JS. NMR in rotating solids.
agation technique for MAS NMR simulation: Applica- J Chem Phys 1979; 70:3300 –3316.
tion to quadrupolar nuclei. J Magn Reson 1998; 132: 28. Varshalovich DA, Moskalev AN, Khersonskii VK.
181–190. Quantum theory of angular momentum. Singapore:
12. Hohwy M, Bildsøe H, Jakobsen HJ, Nielsen NC. Effi- World Scientific; 1988.
cient spectral simulations in NMR of rotating solids. 29. Sakurai JJ. Modern quantum mechanics. New York:
The ␥-COMPUTE algorithm. J Magn Reson 1999; 136: Addison–Wesley; 1994.
6 –14. 30. Press WH, Flannery BP, Teukolsky SA, Vetterling
13. Hodgkinson P, Emsley L. Numerical simulation of sol- VT. Numerical recipes in C. The art of scientific
id-state NMR experiments. Prog NMR Spectrosc 2000; computing. Cambridge, UK: Cambridge University
36:201–239. Press; 1986.
14. Bak M, Rasmussen JT, Nielsen NC. SIMPSON: 31. Jeener J. Emphasising the role of time in quantum
A general simulation program for solid-state dynamics. Bull Magn Res 1994; 16:35– 42.
NMR spectroscopy. J Magn Reson 2000; 147:296 – 32. Abramowitz M, Stegun IA, editors. Handbook of math-
330. ematical functions with formulas, graphs and mathe-
15. Charpentier T, Fermon C, Virlet J. Numerical and matical tables. New York: Dover; 1972.
theoretical analysis of multiquantum magic-angle- 33. Vega AJ. In: Encyclopedia of NMR. New York: Wiley;
spinning experiments. J Chem Phys 1998; 109:3116 – 1995.
3130. 34. Farrar TC. Density matrices in NMR spectroscopy. II.
16. Brinkmann A, Edén M, Levitt MH. Synchronous Concepts Magn Reson 1990; 2:55– 61.
helical pulse sequences in magic-angle-spinning 35. Levitt MH. The signs of frequencies and phases in
NMR. Double quantum spectroscopy of recoupled NMR. J Magn Reson 1997; 126:164 –182.
multiple-spin systems. J Chem Phys 2000; 112: 36. Antzutkin ON. Sideband manipulation in magic-angle-
8539 – 8554. spinning nuclear magnetic resonance. Prog NMR Spec-
17. Kubo A, Imashiro F, Terao T. Fine structures of 1H- trosc 1999; 35:203–266.
coupled 13C MAS NMR spectra for uniaxially rotating 37. Levitt MH. Why do spinning sidebands have the same
molecules in deuterated surroundings: Conformations phase? J Magn Reson 1989; 82:427– 433.
154 EDÉN

38. Edén M, Levitt MH. Computation of orientational av- BIOGRAPHY


erages in solid-state NMR by Gaussian spherical Mattias Edén was born in 1971 in Stock-
quadrature. J Magn Reson 1998; 132:220 –239. holm, Sweden. He received his B.S. in chem-
39. Schmidt A, Vega S. The Floquet theory of nuclear istry from Stockholm University in 1994 and
magnetic resonance spectroscopy of single spins and continued there with doctoral studies in the
dipolar coupled spin pairs in rotating solids. J Chem group of Malcolm H. Levitt. Their work pri-
Phys 1992; 96:2655–2680. marily involved pulse sequence design for
40. Nakai T, McDowell CA. Application of Floquet theory determining molecular structures by solid-
state NMR, as well as the development of
to the nuclear magnetic resonance spectra of homo-
methods for numerically simulating NMR
nuclear two-spin systems in rotating solids. J Chem
experiments. He did postdoctoral work in the field of solid-state
Phys 1992; 96:3452–3465. NMR on quadrupolar nuclei with Lucio Frydman at the University
41. Luz Z, Poupko R, Alexander S. Theory of dynamic of Illinois at Chicago (2000) and at the Weizmann Institute of
magic angle spinning nuclear magnetic resonance and Science in Israel (2001). Currently, Dr. Edén is an assistant professor
its application to carbon-13 in solid bullvalene. J Chem at Stockholm University, focusing on solid-state NMR methodology
Phys 1993; 99:7544 –7553. development for structural investigations of inorganic materials.

You might also like