Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

FULL PAPER WWW.C-CHEM.

ORG

CERES: An Ab Initio Code Dedicated to the Calculation of


the Electronic Structure and Magnetic Properties of
Lanthanide Complexes
Simone Calvello , Matteo Piccardo , Shashank Vittal Rao, and Alessandro Soncini

We have developed and implemented a new ab initio code, efficient library for the direct evaluation of molecular integrals,
CERES (Computational Emulator of Rare Earth Systems), and problem-specific density matrix guess strategies. After
completely written in C1111, which is dedicated to the effi- describing the main features of the new code, we compare its
cient calculation of the electronic structure and magnetic efficiency with the current state–of–the–art ab initio strategy
properties of the crystal field states arising from the splitting to determine crystal field levels and properties, and show that
of the ground state spin-orbit multiplet in lanthanide com- our methodology, as implemented in CERES, represents a more
plexes. The new code gains efficiency via an optimized imple- time-efficient computational strategy for the evaluation of the
mentation of a direct configurational averaged Hartree–Fock magnetic properties of lanthanide complexes, also allowing a
(CAHF) algorithm for the determination of 4f quasi-atomic full representation of non-perturbative spin-orbit coupling
active orbitals common to all multi-electron spin manifolds effects. V
C 2017 Wiley Periodicals, Inc.

contributing to the ground spin-orbit multiplet of the lantha-


nide ion. The new CAHF implementation is based on quasi- DOI: 10.1002/jcc.25113
Newton convergence acceleration techniques coupled to an

Introduction operator, and obtain the resulting crystal field states (RASSI
step). This method was first used to theoretically explain the
The magnetic properties of lanthanide single–molecule magnets presence of a non–magnetic ground state in a Dy3 triangle,[8]
(SMMs) have made them important targets for a wide range of and has since been used to successfully rationalize the mag-
applications, ranging from magnetic resonance contrast agents[1] netic properties of a range of lanthanide SMMs.[9–24]
to building blocks for molecular magnetic memories.[2] The main Spin–orbit coupling is one of the dominant energy scales in
reason for their versatility lies in the details of their electronic lanthanide complexes so that it would be desirable to use all
structure and, in particular, in the almost complete lack of cova- CASSCF optimized spin states to represent the SOC Hamilto-
lent interactions between lanthanide and ligand orbitals for the nian matrix, but in current CASSCF implementations this
lowest energy states which, in turn, leads to a quasi–atomic approach can be computationally very demanding. In the
valence space constituted by LnIII 4f–like orbitals. Their wave RASSI step, furthermore, all CASSCF spin state wave functions
function, therefore, will be dominated by spin–orbit J–multiplets have to be converted to a common molecular orbital basis.
weakly split by electrostatic interactions with the ligand crystal The latter process can also be computationally demanding,
field, causing a strong magnetic anisotropy which is fundamental and even in efficient programs like MOLCAS the maximum num-
for their multiple applications.[2,3] ber of CASSCF spin states which can be currently handled is
Ab initio calculations have proven to be very useful in around 300,[23] thus not allowing a full representation of the
describing the magnetic properties and identifying magneto- SOC interaction in many experimentally relevant ions like TbIII ;
structural correlations for a number of lanthanide complexes DyIII and HoIII .
displaying SMM behavior, thus becoming a common feature in In a recent work,[25] we have discussed the limits of the
experimental studies in the literature. The method of choice CASSCF/RASSI–SO approach for lanthanide SMMs and pro-
for such studies is a combination of Complete Active Space posed an alternative computational strategy, called CAHF/
Self–Consistent Field[4,5] and Restricted Active Space State CASCI–SO (Configurational Average Hartree–Fock/Complete
Interaction via Spin–Orbit coupling[6] (CASSCF/RASSI–SO), as Active Space Configuration Interaction with Spin–Orbit Cou-
for instance implemented in the quantum chemistry program pling), which allows the calculation of magnetic properties of
MOLCAS.[7] Given a mononuclear LnIII complex with an open–
shell configuration of 4f n, out of all multiconfigurational spin S. Calvello, M. Piccardo, S. V. Rao, A. Soncini
states arising from the distribution of n valence electrons in School of Chemistry, University of Melbourne, Victoria, 3010, Australia
E-mail:asoncini@unimelb.edu.au
the seven 4f–like valence orbitals, a subset is chosen and, in
Contract grant sponsor: Australian Research Council (Discovery Project);
the CASSCF step of the calculation, their wave functions are Contract grant number: DP150103254; Contract grant sponsor: Australian
optimized. Next, the spin-free CASSCF wave functions are used Government Research Training Program Scholarship (to S.C. and S.V.R.)
as basis states to diagonalize the spin-orbit coupling (SOC) C 2017 Wiley Periodicals, Inc.
V

328 Journal of Computational Chemistry 2018, 39, 328–337 WWW.CHEMISTRYVIEWS.COM


WWW.C-CHEM.ORG FULL PAPER

LnIII complexes where all spin states are included. In the pre- single Slater determinants (CASCI-SO step), and diagonalize it
liminary calculations shown there, the new method did not to obtain the crystal field levels. The CAHF/CASCI-SO strategy
introduce any significant deviation from CASSCF/RASSI–SO has now been implemented in an efficient C1111 code,
crystal field levels and magnetic properties. In this article, we CERES, which is described in some more detail in what
present a dedicated program for CAHF/CASCI–SO calculations, follows.
called Computational Emulator of Rare Earth Systems In the rest of the Section we will provide a formulation of
(CERES),[26] showcasing its main features and comparing its the CAHF procedure using second quantization, which is
efficiency with the more established CASSCF/RASSI–SO com- known to be more amenable to the definition of efficient con-
putational strategy for the crystal field states and magnetic vergence algorithms. From this point on, therefore, we will
properties of LnIII complexes, as implemented in the software consider a general open–shell lanthanide complex, and we will
MOLCAS. partition its orbitals into three subspaces: mI inactive orbitals,
always doubly occupied, mA active orbitals, occupied by nA
CAHF/CASCI–SO Theory as Implemented in electrons, and mV virtual orbitals, always unoccupied. We
CERES will then define average orbital space occupation numbers as
mX 5 mnXX for the subspace X. Finally, molecular orbital indices
The current ab initio strategy to determine the crystal field
will be partitioned as follows: i; j; k; l for inactive orbitals, u; v;
states in LnIII complexes consists of performing a set of State
w; x for active orbitals, a; b; c; d for virtual orbitals, and p; q; r; s
Averaged CASSCF (SA-CASSCF) calculations.[23] In particular,
for unspecified orbitals.
for each spin quantum number allowed by the occupation of
The well-known second quantization formulation of Self-
the 4f active space, the SA-CASSCF energy functional used in
Consistent Field theory[30] is briefly rehearsed below, to expose
the orbital optimization problem is defined as the average
its specificities when applied to the configurationally averaged
energy of all multiconfigurational electronic eigenstates with
problem implemented in the CAHF module of our code CERES.
a given spin. The SOC operator is then represented on the
An optimal parameterization of the Born-Oppenheimer elec-
basis of as many SA-CASSCF eigenstates as it is computation-
tronic energy functional, leading to a formulation of CAHF as
ally feasible to include, and then diagonalized (RASSI-SO
an unconstrained energy minimization problem, is achieved by
step), yielding the crystal field levels.[23] Note that SA-CASSCF
imposing an exponential parameterization of the (unitary)
eigenstates with different spins feature different (non-orthog-
operator performing orbital rotations in the multi-electron
onal) sets of optimized molecular orbitals for different spin
wavefunction. For instance, given the SD jUi ðjÞi wavefunction,
manifolds, which can make the calculation inefficient. Note
we have:
also that while in a general SA-CASSCF calculation orbital
rotations are coupled to the CI coefficients optimization prob-
jUi ðjÞi5e2^j jUi i (1)
lem, when averaging is carried out over all possible multicon-
figurational eigenstates within a given active space carrying a X ! "
^5
with j jpq Epq 2Eqp , where, according to the usual nota-
specific irreducible representation (irrep) of a given symmetry p>q
group (e.g., having a given total spin), the average energy tion, if apr is an annihilation operator destroying an electron
becomes the trace of the electrostatic Hamiltonian matrix with spin r in the molecular orbital /p , the singlet excitation
within a given irrep of the symmetry group, which, being X †
invariant under a unitary transformation, becomes indepen- operator reads Epq 5 apr aqr . The argument of the SD wave-
r
dent of the eigenstates of the problem. In other words, the function jUi ðjÞi is the antisymmetric matrix j, containing all
orbital optimization problem becomes exactly decoupled non–redundant orbital rotation parameters, that is, rotations
from the CI problem in typical applications of SA-CASSCF to between different orbital subspaces (inactive, active, and vir-
LnIII complexes, and could in principle be solved as a single tual) defined in the CAHF calculation.
CAHF problem.[25,27–29] On account of strong SOC, and considering that in our tar-
This simple observation is the starting point for our pro-
get systems the active space spans the 4f atomic angular
posed CAHF/CASCI-SO strategy. In particular, on account of
momentum shell well-shielded from covalency effects by 5s
the fact that strong-spin orbit coupling does not conserve
and 5p electrons, especially for the smaller lanthanide ions of
the molecule’s spin angular momentum, in the proposed
interest to molecular magnetism due to their large ground
CAHF/CASCI-SO we relax the condition that state-averaging
state angular momentum, we assume it is physically reason-
should be carried out within a given spin manifold. Hence,
able to fully relax the spin-symmetry of the system, and define
the average energy functional is simply the trace of the CI
the CAHF energy functional as the average energy of all the
matrix represented on the basis of all possible Slater deter-
SDs jUi ðjÞi; i51; . . . ; NSD , spanning the chosen 4f active space
minants (SDs), of any MS quantum number, spanning the !
14
chosen active space. After having thus optimized average 4f (i.e., for single-ion lanthanide complexes, NSD 5 , where n
orbitals common to all multiconfigurational spin states aris- n
ing from a given occupation of the 4f active space (CAHF is the number of 4f electrons in the chosen lanthanide ion),
step), we use these orbitals to build a representation of the irrespective of their MS quantum number. This leads to the fol-
sum of electrostatic and SOC Hamiltionians on the basis of lowing expression for the energy:

Journal of Computational Chemistry 2018, 39, 328–337 329


FULL PAPER WWW.C-CHEM.ORG

X
NSD @ 2 E!
E!ðjÞ5N21 hUi ðjÞjHel jUi ðjÞi @jpq @jrs jj50 is the configurationally averaged molecular Hessian
SD
ð1Þ@E !
i
(2) and E pq 5@j j
pq j50
is the configurationally averaged molecular
X 1X
5 hpq D! pq 1 gpqrs d!pqrs 1Enuc gradient, which can be shown to read:
pq
2 pqrs
% &
ð1Þ ð1Þ ð2Þ
E!iu 52 mI F!iu 2mA F!ui (8a)
where Hel is the molecular electrostatic Hamiltonian, hpq and
gpqrs are the mono– and bi–electronic molecular integrals in ð1Þ ð1Þ
E!ia 54F!ia (8b)
the molecular orbital basis expressed in Mulliken notation, and
ð1Þ ð2Þ
the mono– (D! pq ) and bi–electronic (d!pqrs ) average density E!ua 52mA F!ua (8c)
matrices in the molecular orbital basis, defined as:
with
X
NSD X

D! pq 5N21
SD hUi japr aqr j Ui i X# $
ð1Þ 1
i r F!pq 5 hqp 1mI gqpii 2 gqiip
2
X
NSD X i
(9)
d!pqrs 5N21
† †
hUi japr arr0 asr0 aqr jUi i; (3) X # $
SD 1
i r;r0 1mA gqpuu 2 gquup
u
2
are here explicitly calculated as:
8 and
>
> mI dpq p inactive
< X# $
!
D pq 5 mA dpq p active (4) ð2Þ 1
> F!pq 5 hqp 1mI gqpii 2 gqiip
>
: i
2
0 p virtual # $ (10)
X 1
8 ! " 1mA k gqpuu 2 gquup
> mI 2dpq drs 2dps drq p; r inactive u
2
>
>
>
> # $
>
> 1
>
> mI mA dpq drs 2 dps drq p inactive; r active The expression for the configurationally averaged molecular
>
>
>
> 2 ð2Þ
< Hessian E is reported in the Supporting Information, along
d!pqrs 5 or (5)
ð1Þ
with a more thorough derivation of the gradient E . Since
>
>
>
> p active; r inactive the computation of Hessian matrix elements is extremely
>
>
>
> ! " time–consuming, a direct solution of eq. (7), either by invert-
>
> mA k 2dpq drs 2dps drq p; r active
>
>
>
: ing the Hessian or by solving iteratively the linear system of
0 otherwise equations, is rarely used. A significant speed–up is provided by
the well–known Quasi–Newton method,[31] based on the
nA 21
with k5 2m A 21
being the average active space electron–elec- secant condition, in which an approximate Hessian is used to
tron repulsion weight. solve eq. (7), and at every iteration the knowledge provided
Substitution of eqs. (4) and (5) into the average energy by the gradients is used to improve such approximation, with
expression eq. (2) gives the CAHF energy expression[25,27]: the most used update scheme being the Broyden–Fletcher–
" Goldfarb–Shanno (BFGS) algorithm[32–35]:
X X# $#
! I 1 % † &% † &
E5m hii 1 giijj 2 gijji †
i j
2 sk yk 1yk Hk yk sk sk
Hk11 5Hk 1 ! † "2
X# 1
$ sk yk
(11)
1mI mA giiuu 2 giuui (6) † †
i;u
2 Hk yk sk 1sk yk Hk
2 †
" s k yk
X X# $#
1
1mA huu 1k guuvv 2 guvvu 1Enuc with Hk being the approximated inverted Hessian, yk 5Ek11 2
ð1Þ
u v
2
ð1Þ
Ek and sk 5jk11 2jk at iteration k.
Minimization of eq. (6) with respect to the matrix of rotation Finally, Hel 1HSOC , where HSOC is an appropriate representa-
parameters leads to the usual linear system of equations tion of the SOC Hamiltonian (vide infra), is represented on the
whose solution defines the family of convergence algorithms basis of the NSD SDs built with the CAHF–optimized molecular
known as second–order methods: orbitals. Hel 1HSOC is then diagonalized to yield the crystal field
energies and multiconfigurational wavefunctions. Note that,
ð2Þ ð1Þ while Hel is block diagonal in the full SD basis, that is, it has
E j52E (7)
matrix elements only between SDs with like MS quantum num-
where the antisymmetric matrix j collecting the orbital rota- bers, HSOC features first-rank irreducible tensor spin operators
ð2Þ coupling a given MS SD subspace, to the two SD subspaces
tions is here arranged in a column vector, and now E pq;rs 5 characterized by MS 61 quantum numbers. Given the basis of

330 Journal of Computational Chemistry 2018, 39, 328–337 WWW.CHEMISTRYVIEWS.COM


WWW.C-CHEM.ORG FULL PAPER

Figure 1. Flow chart of CERES.

SDs which, for a given MS quantum number, can be naturally the output data. All matrix operations are performed using the
represented as strings of a and b electrons partially occupying fast and efficient C11 EIGEN template library for linear alge-
the active space molecular orbitals, the CI routine is naturally bra.[39] A particular attention has been dedicated to the paral-
based on the r–algorithm developed by Olsen,[30,36] with the lelization of the code where appropriate, by the use of
addition of spin non-conserving single excitations to represent openMP application programming interface (API) specification
the one-electron SOC Hamiltonian: for parallel programming.
X The structure of the implemented CAHF/CASCI-SO algorithm
HSOC 5 lðiÞ # sðiÞ; (12) is sketched in Figure 1. The CAHF function consists of a direct
i
SCF module, where the integral calculation at each iteration
always occurs on the fly, using the LIBINT library for high-
where the summation is over all electrons. In second quantiza-
performance Gaussian integrals computation,[40] which is
tion, this gives
based on the Obara-Saika recursion scheme. A version of the
1X z † † code which is currently being debugged is including the more
HSOC 5 lpq ðapa aqa 2apb aqb Þ general and efficient LIBCINT library for Gaussian basis func-
2 pq
(13)
tions,[41] based on the Rys-polynomials scheme, to overcome
1 † 2 †
1lpq apb aqa 1lpq apa aqb some LIBINT limitations in handling multiple contractions, and
to implement all the relativistic and spin-orbit integrals, which
z 6 x y
where lpq and lpq 5lpq 6ilpq are the integrals of the (bare or are of crucial importance to accurately describe lanthanide
mean–field) one–electron SOC Hamiltonian in the active complexes. All direct CAHF calculations presented here have
molecular orbital basis representation. Generalization of the been performed with the LIBINT library, as this implementation
algorithm to include two–electron SOC contributions is has been extensively debugged. At this stage, the relativistic
underway.[37] integrals entering both the scalar Douglas–Kroll–Hess (DKH)
corrections to the one-electron Hamiltonian[42] and the one-
electron SOC Hamiltonian have been calculated by the use of
CERES Structure
an ad hoc modified version of the LIBINT C11 library. For the
In this section, we present CERES,[26] a quantum chemistry code purpose of comparing our numerical results with those
specifically designed for efficient CAHF/CASCI-SO calculations obtained via the CASSCF/RASSI–SO strategy as implemented in
of crystal field states and magnetic properties of lanthanide MOLCAS, we decided to read in our CASCI module the SOC
complexes. After discussing the CERES main features, we test its atomic mean–field integrals (AMFI) calculated with MOLCAS in
performances and efficiency with respect to the so far the calculations presented here, postponing the comparison
adopted SA-CASSCF/RASSI-SO strategy. between different approximations of the SOC interactions to a
The program CERES is written in C1111, taking advantage of separate forthcoming study.[37]
object-oriented programming. Implemented by making use of CERES features several algorithms for the estimation of the
the open-source C11 BOOST libraries,[38] a PYTHON front-end initial orbital guess feeding the CAHF iterative calculation,
makes CERES user-friendly, and automates many common tasks some of which are well known, namely the diagonalization of
such as parsing the input data, building jobs as a sequence of the mono–electronic Hamiltonian and the projection on a
desired methods, the re–execution of a sequence of calcula- larger basis set of the density matrices obtained from a CAHF
tions on a set of molecules, or a quick and direct control of calculation on a smaller basis set. As previous studies indicated

Journal of Computational Chemistry 2018, 39, 328–337 331


FULL PAPER WWW.C-CHEM.ORG

that simple charge models are quite effective in predicting the atomic orbital basis. Convergence criterions are the same used
ð1Þ
properties of lanthanide complexes,[21] however, we chose to in MOLCAS,[7] namely DE! and jjE! jj1 .
implement a generalized form of the Sum of Atomic Densities CERES features a mixed first–second order convergence algo-
guess[43] specifically designed for calculations of LnIII SMM rithm, using DIIS[48,49] with the molecular gradient as the error
magnetic properties, in which we assume that the initial choice[50] in the first part. Once sufficiently close to conver-
ð1Þ ! !
atomic orbitals are orthogonal and localized. The resulting gence, using jjE! jj1 and jjDRjj 1 as criterions, with R being
density matrices will thus be diagonal, with their non–zero ele- the average density matrix in the atomic orbital basis, the sec-
ments corresponding to the occupation number of the atomic ond–order convergence algorithm is used. CERES implements
n
orbital whose angular momentum is l, computed as mn;l 5 4l12 . both an iterative solution of eq. (7) via the Jacobi method and
Chemical knowledge of the target molecule can be translated a Limited–Memory BFGS Quasi–Newton algorithm,[51] whose
into input partial charges to be distributed on selected atoms main advantage with respect to the BFGS algorithm is the pos-
or ligands. Density matrices generated with this procedure are, sibility to compute the updated Hessian without storing it in
then, converted into the non–orthogonal basis using an memory, thus improving the scalability of the method.
€wdin transformation:
inverse Lo Although most QN algorithms use identity matrix as initial
Hessian approximation, we have chosen to use the one–elec-
1 1
RXnon2orth 5S22 RXorth S22 (14) tron component of the diagonal elements of the exact molec-
ular Hessian so as to improve the efficiency of QN and avoid
where S is the overlap matrix in the atomic orbital basis. CERES time–consuming line–search methods.[52]
also allows to create such diagonal density matrices in a Once CAHF convergence has been achieved, the optimized
smaller basis set and project them in the final basis set via a LCAO coefficient matrix is used to set up the CASCI–SO matrix.
mixed–basis bi–electronic integral contraction. In the current version of CERES, spin–orbit contributions are
The direct SCF algorithm we have implemented[44] features either based on the LIBINT library, either using the Breit-Pauli
a preliminary screening of the list of overlap distributions for a Hamiltonian, soon to be interfaced with the Cholesky decom-
given basis set, achieved by computing the overlap between position of the bi–electronic spin–orbit integrals,[37] or by
basis set shells, and by avoiding computation of all integrals using the well–known AMFI approximation[53] in various for-
between such shells if, provided they do not belong to the mulations, including the possibility to read them from files
same atom, the Euclidean norm of their overlap is lower than produced by MOLCAS, as done here for ease of comparison. The
a chosen threshold. Second, at each iteration a Cauchy– optimized LCAO coefficient matrix is, then, used to transform
Schwarz screening procedure is used,[45] with a dynamical electron repulsion bi–electronic integrals into the molecular
screening threshold defined so as to improve the precision of orbital active subspace via two successive two–index semi–
the calculation closer to convergence: transformations.[54] Cholesky decomposition of the electron
# $ !! repulsion integrals is also going to be soon used in this step,
ð1Þ
1 29 jjE jj1 so that the integrals on the molecular basis are going to be
d5min min ; 10 ; max ;e (15)
Scond 1027 re-composed as a sum of Cholesky vectors transformed on the
molecular basis. However, results presented here do not yet
where Scond is the condition number of the overlap matrix, ! is make use of this facility. The electron-repulsion and SOC inte-
the machine precision, and jjxjj1 5maxðjx1 j; jx2 j; # # #Þ. Finally, grals transformed on the molecular basis are finally used to
CERES uses an incremental Fock build algorithm,[46] which build the matrix for the CASCI–SO problem, the implementa-
becomes particularly efficient when coupled with Cauchy– tion of which is based on the r–algorithm developed by
Schwarz screening because the difference between consecu- Olsen.[36]
tive iterations density matrices is, in general, smaller when
closer to convergence. Incremental Fock build is started when Analysis of CERES Performance
sufficiently close to convergence and, as it introduces errors in
the energy, it is reset every eight iterations by performing a Although CAHF/CASCI–SO and SA–CASSCF/RASSI–SO are two
non–incremental iteration. Precision of the final result is conceptually different methods that cannot be directly com-
granted by turning off incremental Fock build when all conver- pared in terms of the speed of each single step in the process,
gence criterions are smaller than an order of magnitude higher both are expected to lead to results of comparable accuracy,
than the convergence thresholds. The effective Fock Hamilto- and an efficient implementation of our proposed CAHF/
nian, which is diagonalized in every first–order iteration,[15,28] CASCI–SO is expected to be competitive with the SA–CASSCF/
includes level shifters k1 and k2[47] so as to separate the ener- RASSI–SO approach for the calculation of the electronic struc-
gies of orbital subspaces and improve the efficiency of the cal- ture and magnetic properties of the crystal field levels arising
culation, which are defined as: from the ground spin–orbit term of any LnIII complex. To pro-
vide evidence of this, and to test the efficiency of our imple-
F ðk1 ; k2 Þ5F1k1 R2 1ðk1 1k2 ÞR3 (16) mentation of the CAHF/CASCI–SO method, we have performed
single point calculations on a set of ten complexes of
where F is the effective CAHF Hamiltonian and R2 and R3 are, TbIII ; DyIII ; HoIII , and ErIII with CERES, comparing the perfor-
respectively, the active and virtual density matrices in the mance with the CASSCF/RASSI–SO strategy as implemented in

332 Journal of Computational Chemistry 2018, 39, 328–337 WWW.CHEMISTRYVIEWS.COM


WWW.C-CHEM.ORG FULL PAPER

Figure 2. Plots of the timings, expressed as hours, for the crystal field level calculations of the chosen complexes by SA–CASSCF/RASSI-SO and CAHF/
CASCI–SO. The four plots represent, respectively, the timings for Tb (top left), Dy (top right), Ho (bottom left), and Er (bottom right). Ligand number on
the x axis is extracted from Table 3 in the Supporting Information. [Color figure can be viewed at wileyonlinelibrary.com]

MOLCAS 8.0.[7] Experimental geometries were available for all Dy convergence criterions set to the same values of MOLCAS as DE
compounds[55–61] and, whenever possible, experimental geom- % 1028 and jjE ð1Þ jj1 % 1024 .
etries have been used for other ions.[55,58,59] All calculations for To discuss both the efficiency and accuracy of the CAHF/
each ion have been performed on separate NeCTAR research CASCI–SO method we will analyze, respectively, the timings
cloud virtual machines,[62] each with 16 GB RAM and 2.3 GHz required for the magnetic properties calculation and the val-
Intel CPUs. ues of crystal field energies, g–tensors and orientation of the
The active space is made of the seven LnIII 4f orbitals occu- main anisotropy axes for the ground state spin–orbit multiplet.
pied by, respectively, 8 electrons in Tb, 9 electrons in Dy, 10 For odd–electron systems, g–tensors and anisotropy axes will
electrons in Ho, and 11 electrons in Er. Given the inability of be computed between degenerate states of each Kramers
LIBINT to efficiently handle general contraction basis sets, we doublet, while for even–electron systems they will be com-
have chosen segmented basis sets, namely SARC2–QZVP–DKH puted between two non–degenerate, consecutive states. We
for LnIII,[63] Ahlrichs–PVDZ for coordinating atoms and Ahl- have reported the computational timings for the overall calcu-
richs–VDZ for all other atoms.[64] Throughout the article, the lations with the SA–CASSCF/RASSI-SO and CAHF/CASCI–SO
following abbreviations will be in use to shorten the names of methods in Figure 2, with the exact timings reported in Table
the molecules: acac 5 acetylacetonate, dppz 5 dipyridophena- 3 and a breakdown into the timings for each phase of all cal-
zine, dpq 5 dipyridoquinoxaline, phen 5 1,10-phenanthroline, culations in Table 4 of the Supporting Information. All values
hfac 5 hexafluoroacetylacetone, dme 5 dimethoxyethane, of crystal field energies and g–tensors for the ground state
paaH$ 5 N-(2-Pyridyl)acetoacetamide, tfpb 5 4,4,4-trifluoro-1- spin–orbit term of each molecule have also been included in
phenyl-1,3-butandionate, tta 5 tetradecylthioacetate, bipy 5 the Supporting Information. To analyze and discuss the accu-
2,20-bipyridine, pinene2bipy 5 4,5-pinene bipyridine. racy of our method with respect to SA–CASSCF/RASSI-SO,
MOLCAS calculations are performed using High Cholesky then, we have computed the relative difference between crys-
option, with a cut–off threshold of 1028. We have performed tal field energies and between the largest components of the
one SA–RASSCF calculation for each possible total spin quan- g–tensors as, respectively,
tum number S by averaging over all possible spin states
except for Tb triplets, for which the number of states has jeRASSI 2eCASCI j
De% 5 # 100
been reduced so as not to exceed virtual machine memory. maxðeRASSI ; eCASCI Þ
The total number of SA–CASSCF calculations performed for jgRASSI 2gCASCI j
each lanthanide complex is, respectively, four for Tb, three for Dg% 5 # 100
maxðgRASSI ; gCASCI Þ
Dy and Ho, and two for Er. In the RASSI step, we have selected
all states from some of the lowest Russell–Saunders terms for where the subscripts RASSI (CASCI) represent a shorthand
each S while still maintaining a total number of states less notation for SA–CASSCF/RASSI–SO (CAHF/CASCI–SO).
than 300.[23] An overview of the number of states used in We have also computed the angle between the main anisot-
each step is presented in the Supporting Information. ropy axes computed with the two methods. We have, then,
CERES calculations include a CAHF calculation on all spin collected data subsets into several plots to highlight the most
states with level shifters for active and virtual space of 0.4 and important conclusions from this study. We have included the

Journal of Computational Chemistry 2018, 39, 328–337 333


FULL PAPER WWW.C-CHEM.ORG

Figure 3. Plots for the accuracy study of the SA–CASSCF/RASSI-SO, and CAHF/CASCI–SO methods. The three plots showcase the accuracy of, respectively,
crystal field levels (left), g–tensors (middle), and anisotropy axes (right). Ligand number on the x axis is extracted from Table 3 in the Supporting Informa-
tion. [Color figure can be viewed at wileyonlinelibrary.com]

highest errors for crystal field levels, g–tensors and anisotropy Despite introducing a significative speed–up, the CAHF/
axes in Figure 3, the relative errors for the g–tensor for the CASCI–SO method does not introduce significant losses in
ground crystal field state and the first two excited states for all accuracy. Figure 3 does, in fact, show remarkable agreement
the 40 molecules in Figure 4 and the relative errors for the for the crystal field levels between the two methods, with an
crystal field energy of the second excited state, which corre- average deviation of De% 51:5% and the highest error being
sponds to a state in the first excited Kramers doublet for odd– 2.47% for ½ErðtfpbÞ3 ðdppzÞ'. Analysis of g–tensors also displays
electron systems and to the lower energy state in the first a good agreement between methods, with the lowest errors
excited pseudo–Kramers doublet for even–electron systems, in arising for odd–electron systems. Even–electron systems show
Figure 5. slightly higher errors, with three notable outliers in ½TbðacacÞ3
The main conclusion which can be drawn from the analysis ðH2 OÞ2 '; ½TbðtfpbÞ3 ðdppzÞ' and ½HoðttaÞ3 ðpinene2bipyÞ'. Ani-
of Figure 2 is that our implementation of the CAHF/CASCI–SO sotropy axes are also reproduced very well by the CAHF/
algorithm is indeed more efficient than SA–CASSCF/RASSI–SO. CASCI–SO method, with the maximum error being lower than
Out of the 40 molecules considered, in fact, the latter 5% for all molecules but one.
displays similar timings only for ½ErðacacÞ3 ðH2 OÞ2 ' and It has to be noted that the most relevant deviations
½ErðacacÞ3 ðphenÞ'. between values usually occur for high energy crystal field

Figure 4. Plots of the relative error for the g–tensor computation for the ground state and the first two excited Kramers doublets, for odd–electron sys-
tems, or Ising doublets, for even–electron systems, between the SA–CASSCF/RASSI-SO, and CAHF/CASCI–SO methods. The four plots represent, respectively,
the g–tensors for Tb (top left), Dy (top right), Ho (bottom left), and Er (bottom right). Ligand number on the x axis is extracted from Table 3 in the Sup-
porting Information. [Color figure can be viewed at wileyonlinelibrary.com]

334 Journal of Computational Chemistry 2018, 39, 328–337 WWW.CHEMISTRYVIEWS.COM


WWW.C-CHEM.ORG FULL PAPER

Figure 5. Plots of the relative error of the energy of the first excited
Kramers doublet, for odd–electron systems, or of the lower energy state of
the first excited Ising doublet, for even–electron systems, between the SA–
CASSCF/RASSI-SO, and CAHF/CASCI–SO methods. Ligand number on the x
axis is extracted from Table 3 in the Supporting Information. [Color figure
Figure 6. Orientation of the main magnetic axes, computed with CERES, of
can be viewed at wileyonlinelibrary.com]
the ground state and of the two highest energy KDs for ½DyðacacÞ3 ðdppzÞ'.
The anisotropy axis of the ground and the two excited KDs are repre-
states, which is confirmed by the plots in Figure 4. When only sented, respectively, in green, red, and blue. [Color figure can be viewed at
the lowest three Kramers or Ising doublets are considered, wileyonlinelibrary.com]
deviations between methods are greatly reduced, with relative
errors usually lower than 2% but for ½TbðtfpbÞ3 ðdppzÞ'. As the second highest–energy KD is almost perpendicular to that
these states are the most significant for explaining several of the ground state MJ 56 15 2 atomic state, while that of the
magnetic phenomena, we can thus safely claim that the highest–energy KD lies at an angle of approximately 458 from
CAHF/CASCI–SO method will be accurate in describing them. the plane containing the other two principal axes. This is illus-
Finally, we note that the energy of the lowest excited crystal trated in Figure 6, where we have plotted the main magnetic
field state is also well reproduced, as shown in Figure 5, with anisotropy axes for the ground state and the two highest
deviations averaging 1%. Overall, these results indicate that energy KDs for ½DyðacacÞ3 ðdppzÞ'. The strong axiality of the
the CAHF/CASCI–SO method, as implemented in CERES, pro- magnetic anisotropy of the ground state KD, whose g–tensor
vides an efficient alternative to the more established CASSCF/ values can be explained by the KD being composed from
RASSI–SO method with the approximations introduced not almost pure MJ 56 15 2 states, can be rationalized via the elec-
leading to big errors. trostatics arguments of Rinehart and Long,[65] based on the
An example of the application of the CAHF/CASCI–SO aspherical electron density distributions of the MJ levels of the
method to the characterization of the magnetic features of lanthanide ions,[66] according to which the lowest–energy KD
lanthanide complexes is obtained by analyzing the relative ori- of the Dy complexes, having an oblate spheroid electronic dis-
entation of the main magnetic anisotropy axes for different tribution, minimizes the electrostatic interaction energy with
KDs of seven of the ten Dy complexes studied here. In Table 1, the ligands charge distribution, determining the direction of
we have reported the highest principal value of the g–tensor the ground state magnetic easy–axis.[21,59] It is interesting to
for the ground KD, and for the two highest energy KDs for the note, however, that both high–energy KDs also possess an
seven complexes of Dy, which we found to have all the follow- almost pure MJ 56 15 2 atomic state character hence a large
ing three properties: (i) they are almost purely of the easy–axis magnetic moment, which is oriented perpendicularly to that
type, (ii) they have a g–tensor principal value that corresponds of the ground state for the second highest–energy KD, as if
to an almost pure MJ 56 15 2 atomic state, with the other two
these energy levels represented transition states for the reori-
principal values being close to zero, (iii) the principal axis of entation of the large (hence almost–classical) Dy magnetic

Table 1. Principal values of the g–tensors for the ground and the two highest energy KDs and the angles between the three main magnetic anisotropy
axes for the seven compounds of Dy.

Molecule g1zz g7zz !7 (cm21) g8zz !8 (cm21) Angle17 (8) Angle18 (8) Angle78 (8) Angle137#8 (8)
½DyðacacÞ3 ðdppzÞ' 19.399 18.666 378 19.775 499 94 49 69 130
½DyðacacÞ3 ðdpqÞ' 19.320 19.548 377 19.852 559 98 51 61 122
½DyðacacÞ3 ðphenÞ' 19.371 18.646 399 19.541 508 91 51 58 125
½DyðhfacÞ3 ðdmeÞ' 19.461 18.836 290 19.799 445 86 48 113 126
½DyðtfpbÞ3 ðdppzÞ' 19.329 19.561 276 19.870 488 91 49 71 132
½DyðttaÞ3 ðbipyÞ' 19.573 18.624 343 19.457 428 97 133 80 136
½DyðttaÞ3 ðphenÞ' 19.530 18.920 339 19.757 512 87 51 68 134
In the table, Angle17 represents the angle between the main magnetic anisotropy axes of the two KDs, while Angle137#8 represents the angle between
the anisotropy axis of KD 8 and the normal to the plane containing the anisotropy axes of KDs 1 and 7

Journal of Computational Chemistry 2018, 39, 328–337 335


FULL PAPER WWW.C-CHEM.ORG

moment. It would be tempting to speculate that the excited [18] M.-E. Boulon, G. Cucinotta, J. Luzon, C. Degl’Innocenti, M. Perfetti, K.
KD (KD 7), with an anisotropy axis almost perfectly perpendic- Bernot, G. Calvez, A. Caneschi, R. Sessoli, Angew. Chem. Int. Ed. 2013,
52, 350.
ular to that of the ground state KD, corresponds to a higher [19] J. Jung, O. Cador, K. Bernot, C. Pointillart, J. Luzon, B. Le Guennic, Beil-
energy stationary point (saddle point) on the classical electro- stein J. Nanotech. 2014, 5, 2267.
statics energy surface, representing an energy barrier for a [20] J. Jung, X. Yi, G. Huang, G. Calvez, C. Daiguebonne, O. Guillou, O.
Cador, A. Caneschi, T. Roisnel, B. L. Guennic, K. Bernot, Dalton Trans.
classical spin–reorientation pathway. The orientation of the
2015, 44, 18270.
magnetic moment of the highest–energy KD seems, instead, [21] N. F. Chilton, D. Collison, E. J. L. McInnes, R. E. P. Winpenny, A. Soncini,
to escape this simple classical rationalization. We limit our- Nat. Commun. 2013, 4, 2551.
selves to observe this interesting pattern resulting from our [22] D. Aravena, E. Ruiz, Inorg. Chem. 2013, 52, 13770.
[23] L. Ungur, L. F. Chibotaru, Lanthanides and Actinides in Molecular Mag-
calculations performed using our new code CERES, as a proper netism; WileyVCH: Weinheim, Germany, 2015; pp. 153–184.
rationalization would require further studies, which go beyond [24] M. Vonci, M. J. Giansiracusa, R. W. Gable, W. Van den Heuvel, K.
the scope of the present article, in which the main focus is the Latham, B. Moubaraki, K. S. Murray, D. Yu, R. A. Mole, A. Soncini, C.
Boskovic, Chem. Commun. 2016, 52, 2091.
discussion of the implementation of the CAHF/CASCI–SO strat-
[25] W. Van den Heuvel, S. Calvello, A. Soncini, Phys. Chem., Chem. Phys.
egy in CERES. 2016, 18, 15807.
[26] A. Soncini, S. Calvello, M. Piccardo, S. V. Rao, Ceres, an ab initio quan-
Keywords: lanthanide single molecule magnets # ab initio # - tum chemistry package for the electronic structure and magnetic prop-
crystal field levels # electronic structure theory # configura- erties of lanthanide complexes, The University of Melbourne, 2017.
[27] R. McWeeny, Mol. Phys. 1974, 28, 1273.
tional average [28] R. McWeeny, Methods of Molecular Quantum Mechanics; Academic
Press: San Diego, 1989.
[29] S. Itoh, R. Saito, T. Kimura, S. Yabushita, J. Phys. Soc. Jpn. 1993, 62, 2924.
[30] T. Helgaker, P. J€ orgensen, J. Olsen, Molecular Electronic–Structure The-
How to cite this article: S. Calvello, M. Piccardo, S. V. Rao, A.
ory; Wiley: New York, 2012.
Soncini, J. Comput. Chem. 2018, 39, 328–337. DOI: 10.1002/ [31] R. Fletcher, Practical Methods of Optimization; Wiley: Chichester, West
jcc.25113 Sussex, 1987.
[32] C. G. Broyden, J. Inst. Math. Appl. 1970, 6, 222.
] Additional Supporting Information may be found in the [33] R. Fletcher, Comput. J. 1970, 13, 317.
online version of this article. [34] D. Goldfarb, Math. Comput. 1970, 109, 23.
[35] D. F. Shanno, Math. Comput. 1970, 111, 647.
[36] J. Olsen, B. O. Roos, P. J€orgensen, H. J. A. Jensen, J. Chem. Phys. 1988,
89, 2185.
[1] M. Bottrill, L. Kwok, N. J. Long, Chem. Soc. Rev. 2006, 35, 557. [37] M. Piccardo, A. Soncini, A full-pivoting algorithm for the Cholesky
[2] D. N. Woodruff, R. A. Layfield, R. E. P. Winpenny, Chem. Rev. 2013, 113, decomposition of two-electron repulsion and spin-orbit coupling inte-
5110. grals. J. Comput. Chem. 2017, 38, 2775.
[3] D. Gatteschi, R. Sessoli, J. Villain, Molecular Nanomagnets; Oxford Uni- [38] B. Schling, The Boost C11 Libraries; XML Press, 2011.
versity Press: New York, 2006. [39] G. Guennebaud, B. Jacob, et al., Eigen v3, 2010. Available at: http://
[4] B. O. Roos, P. R. Taylor, P. E. M. Siegbahn, Chem. Phys. 1980, 48, 157. eigen.tuxfamily.org. Last accessed August 18, 2017.
[5] P. E. M. Siegbahn, J. Alml€ of, A. Heiberg, B. O. Roos, J. Chem. Phys. [40] E. F. Valeev, Libint: A Library for the Evaluation of Molecular Integrals of
1981, 74, 2384. Many-Body Operators over Gaussian Functions, Version 2.2, Modified by
[6] P.-R. Malmqvist, B. O. Roos, B. Schimmelpfennig, Chem. Phys. Lett. S. V. Rao, M. Piccardo, A. Soncini, Libint, 2016. Available at: http://
2002, 357, 230. libint.valeyev.net/. Last accessed August 18, 2017.
[7] F. Aquilante, J. Autschbach, R. K. Carlson, L. F. Chibotaru, M. G. Delcey, [41] Q. Sun, J. Comput. Chem. 2015, 36, 1664.
L. De Vico, I. F. Galv#an, N. Ferr#e, L. M. Frutos, L. Gagliardi, M. Garavelli, [42] M. Douglas, N. M. Kroll, Ann. Phys. 1974, 82, 89.
A. Giussani, C. E. Hoyer, G. Li Manni, H. Lischka, D. Ma, P. r. Malmqvist, [43] J. H. Van Lenthe, R. Zwaans, H. J. J. Van Dam, M. F. Guest, J. Comput.
T. Muller, A. Nenov, M. Olivucci, T. B. Pedersen, D. Peng, F. Plasser, B. Chem. 2006, 27, 926.
Pritchard, M. Reiher, I. Rivalta, I. Schapiro, J. Segarra-Mart#ı, M. Stenrup, [44] J. Alml€of, K. Faegri, K. Korsell, J. Comput. Chem. 1982, 3, 385.
D. G. Truhlar, L. Ungur, A. Valentini, S. Vancoillie, V. Veryazov, V. P. [45] M. H€aser, R. Ahlrichs, J. Comput. Chem. 1989, 10, 104.
Vysotskiy, O. Weingart, F. Zapata, R. Lindh, J. Comput. Chem. 2016, 5, [46] E. Schwegler, M. Challacombe, M. Head-Gordon, J. Phys. Chem. 1997,
506. 106, 9708.
[8] L. F. Chibotaru, L. Ungur, A. Soncini, Angew. Chem. Int. Ed. 2008, 47, [47] M. F. Guest, V. R. Saunders, Mol. Phys. 1974, 28, 819.
4126. [48] P. Pulay, Chem. Phys. Lett. 1980, 73, 393.
[9] L. F. Chibotaru, L. Ungur, J. Chem. Phys. 2012, 137, 064112. [49] P. Pulay, J. Comp. Chem. 1982, 3, 556.
[10] L. Ungur, L. F. Chibotaru, Chem. Eur. J. 2017, 23, 3708. [50] V. I. Ionova, E. A. Carter, J. Comput. Chem. (1996), 17, 1836.
[11] L. Ungur, L. F. Chibotaru, Phys. Chem. Chem. Phys. 2011, 13, 20086. [51] J. Nocedal, Math. Comput. 1980, 151, 773.
[12] R. Marx, F. Moro, M. D€ orfel, L. Ungur, M. Waters, S. D. Jiang, M. Orlita, [52] L. Armijo, Pac. J. Math. 1966, 16, 1.
J. Taylor, W. Frey, L. F. Chibotaru, J. van Slageren, Chem. Sci. 2014, 5, [53] A. B. Heb, C. M. Mariana, U. Wahlgrenb, O. Gropenc, Chem. Phys. Lett.
3287. 1996, 251, 365.
[13] M. E. Boulon, G. Cucinotta, S. S. Liu, S. D. Jiang, L. Ungur, L. F. [54] S. Wilson, Methods in Computational Chemistry; Plenum Press: New
Chibotaru, S. Gao, R. Sessoli, Chem.–Eur. J. 2013, 19, 13726. York, 1987; pp. 251–309.
[14] J. Luzon, R. Sessoli, Dalton Trans. 2012, 41, 13556. [55] S. D. Jiang, B. W. Wang, G. Su, Z. M. Wang, S. Gao, Angew. Chem. Int.
[15] K. Bernot, J. Luzon, A. Caneschi, D. Gatteschi, R. Sessoli, L. Bogani, A. Ed. 2010, 49, 7448.
Vindigni, A. Rettori, M. G. Pini, Phys. Rev. B 2009, 79, 134419. [56] G. J. Chen, C. Y. Gao, J. L. Tian, J. Tang, W. Gu, X. Liu, S. P. Yan, D. Z.
[16] K. Bernot, J. Luzon, L. Bogani, M. Etienne, C. Sangregorio, M. Liao, P. Cheng, Dalton Trans. 2011, 40, 5579.
Shanmugam, A. Caneschi, R. Sessoli, D. Gatteschi, J. Am. Chem. Soc. [57] G. J. Chen, Y. N. Guo, J. L. Tian, J. Tang, W. Gu, X. Liu, S. P. Yan, P.
2009, 131, 5573. Cheng, D. Z. Liao, Chem. Eur. J. 2012, 18, 2484.
[17] G. Cucinotta, M. Perfetti, J. Luzon, M. Etienne, P.-E. Car, A. Caneschi, G. [58] E. M. Fatila, E. E. Hetherington, M. Jennings, A. J. Lough, K. E. Preuss,
Calvez, K. Bernot, R. Sessoli, Angew. Chem. Int. Ed. 2012, 51, 1606. Dalton Trans. 2012, 41, 1352.

336 Journal of Computational Chemistry 2018, 39, 328–337 WWW.CHEMISTRYVIEWS.COM


WWW.C-CHEM.ORG FULL PAPER

[59] N. F. Chilton, S. K. Langley, B. Moubaraki, A. Soncini, S. R. Batten, K. S. [64] A. Sch€afer, H. Horn, R. Ahlrichs, J. Chem. Phys. 1992, 97, 2571.
Murray, Chem. Sci. 2013, 4, 1719. [65] J. Rinehart, J. Long, Chem. Sci. 2011, 2, 2078.
[60] Z. G. Wang, J. Lu, C. Y. Gao, C. Wang, J. L. Tian, W. Gu, X. Liu, S. P. Yan, [66] J. Sievers, Z. Phys. B Condens. Matter 1982, 45, 289.
Inorg. Chem. Commun. 2013, 27, 127.
[61] Y. Bi, Y. N. Guo, L. Zhao, Y. Guo, S. Y. Lin, S. D. Jiang, J. Tang, B. W.
Wang, S. Gao, Chem. Eur. J. 2011, 17, 12476. Received: 18 August 2017
[62] Nectar Research Cloud, Available at: https://nectar.org.au/. Last Revised: 30 October 2017
Accessed October 30, 2017. Accepted: 31 October 2017
[63] D. Aravena, F. Neese, D. A. Pantazis, J. Chem. Theory Comput. 2016, 12, 1148. Published online on 20 November 2017

Journal of Computational Chemistry 2018, 39, 328–337 337

You might also like