CASSCF Tutorial

Download as pdf or txt
Download as pdf or txt
You are on page 1of 90

1

CASSCF Calculations in ORCA (4.2):


A tutorial Introduction
D. Aravena, M. Atanasov, V. G. Chilkuri, Y. Guo, J. Jung, D. Maganas, B. Mondal,
I. Schapiro, K. Sivalingam, S. Ye and F. Neese

Contents
1 Introduction .................................................................................................................................................... 2
2 The [Cr(NH3)6]3+ complex – a standard run .................................................................................... 4
2.1 Organization of the Calculation: Preliminary Chemical Considerations ................... 4
2.2 Initial 3d CAS(3,5)................................................................................................................................ 5
2.3 Including ligand orbitals CAS(7,7) ............................................................................................... 7
2.4 Including a second d-shell CAS(7,12) ......................................................................................... 8
2.5 Reading the wave function ........................................................................................................... 11
2.6 NEVPT2 CAS(7,12) ........................................................................................................................... 12
3 [Cr(NH3)6]3+ complex - extracting ligand field parameters. ................................................. 14
4 [CrCl6]3- model complex - CASSCF for larger active spaces................................................... 18
5 A comment on using ANO basis sets ................................................................................................ 22
6 [FeIV(O)(TMC)(MeCN)]2+ - covalent metal-ligand interactions and the computation
of MCD / Mössbauer .......................................................................................................................................... 24
6.1 Calculation of MCD spectra and Quadrupole splitting .................................................... 30
6.2 Mössbauer Parameters .................................................................................................................. 33
7 [Co(SH)4]2- - Optical and Magnetic properties............................................................................. 34
7.1 Electronic Structure ......................................................................................................................... 34
7.2 Setting up the CASSCF/NEVPT2 calculation ........................................................................ 35
7.3 Optical spectra .................................................................................................................................... 37
7.4 G-tensor ................................................................................................................................................. 38
7.5 D-tensor ................................................................................................................................................. 41
8 Cu-dimer J-Couplings with NEVPT2, DLPNO-NEVPT2 and MRCI ...................................... 42
8.1 CASSCF(2,2) and NEVPT2 ............................................................................................................. 42
8.2 CASSCF(2,2) DDCI3 - the game changer ............................................................................... 44
8.3 RI approximation for CASSCF, NEVPT2, DLPNO-NEVPT2 and MRCI ...................... 46
8.4 CASSCF(18,10) and NEVPT2 ....................................................................................................... 48
8.5 Wave function Printing .................................................................................................................. 50
8.5.1 CASSCF and ICE......................................................................................................................... 50
8.5.2 MRCI ............................................................................................................................................... 50
2

9 A comment about CASSCF calculations on heavier elements (lanthanide- and


actinide-based systems) .................................................................................................................................. 52
9.1 Basis, RI-Basis and ECP .................................................................................................................. 52
9.2 Initial guess .......................................................................................................................................... 52
9.3 Additional features concerning CASSCF calculations on f-elements systems ..... 53
9.4 Example: Investigation of the spectroscopic and magnetic properties of the
Cs2NaDyCl6 elpasolite ................................................................................................................................... 54
10 Fragment Derived Guess (orca_mergefrag) ................................................................................. 60
11 Manipulation of the ORCA GBW File (orbitals)........................................................................... 62
12 p-Phenylenediamine (PPD) ground state - organic molecule ............................................. 65
13 Adenine spectra – using symmetry .................................................................................................. 68
14 NEVPT2-F12 – reaching the basis set limit................................................................................... 72
15 Appendix [Cr(NH3)6]3+ - detailed walkthrough .......................................................................... 74
15.1 [Cr(NH3)6]3+ complex - a walk through tutorial guide................................................ 74
15.2 Organization of the Calculation: Preliminary Chemical Considerations ........... 75
15.3 Initial guess CASSCF .................................................................................................................... 76
15.4 State-Averaging as convergence aid.................................................................................... 80
15.5 Including ligand orbitals ........................................................................................................... 81
15.6 Including a second d-shell........................................................................................................ 82
15.7 Basis set projection (increasing the basis set) ............................................................... 84
15.8 Calculating the ligand field spectrum ................................................................................. 87

1 Introduction
The complete active space self-consistent field (CASSCF) theory is one of the most used
and powerful methods for electronic structure calculations of transition metals. In
contrast to widely used DFT methods, CASSCF is not a black box method because it
requires the user to select a set of orbitals and electrons that constitutes the active
space. Hence, the input for a CASSCF calculation is case-specific and requires some
consideration from the user with respect to the computational setup. This tutorial is
complementary to the section “running typical calculation” of the ORCA manual. The aim
of this tutorial is to introduce possible strategies to carry out such CASSCF calculations
on transition metal complexes.

Here is a list of details to be considered:


· Inclusion of all d-orbitals in the active space
· Inclusion of the second d-shell in the active space
· Addition of a few ligand orbitals in the active space
· State-averaging of excited states and different spin multiplicities
· Use of a large basis set
· Use of an ANO basis set
3

· Interpreting the wave function (CSFs and spin determinants)


· Ab initio ligand field analysis
· Property calculations (MCD, Mössbauer parameters, ZFS, g-tensor, ...)
· Inclusion of dynamic electron correlation (NEVPT2/MRCI)
· RI Approximation in CASSCF, NEVPT2 and MRCI
· Using point-group symmetry in ORCA

Before running a CASSCF calculation of a transition metal it is crucial to have some


chemical intuition and a basic understanding of the electronic structure of the selected
system.

In the following sections, we will go step by step through a few examples and illustrate
how one can approach all these considerations. By doing so, we explore several “initial
guesses”, convergence strategies and interesting features of the CASSCF program. For
more details on the available options and the general program usage we refer to the
ORCA manual.

Before diving into the practical examples, note that a converged CASSCF wave function is
the starting point for a subsequent multireference calculation that takes into account
dynamic electron correlation. Commonly applied multireference methods are
multireference configuration interaction (MRCI) and multireference perturbation theory
(MRPT). The internally contracted N-Electron valence state perturbation theory
(NEVPT2) is often the first choice to include dynamical correlation. 1,2,3 The method is
efficient and requires only one additional keyword to the CASSCF input:
%casscf
...
PTMethod SC_NEVPT2 # for the strongly contracted NEVPT2
# Other options: FIC-NEVPT2, DLPNO-NEVPT2, FIC_CASPT2
end

In previous version of ORCA, “NEVPT2” abbreviates the “strongly contracted” version of


the NEVPT2 approach. Starting with ORCA 4.0, the PC-NEVPT2 approach in the
canonical and in the DLPNO methodology is available.4 Since the employed wave
function is fully internally contracted (FIC), we prefer to call the method FIC-NEVPT2.
Details and input examples are discussed in Section 8. There we will also illustrate
calculations with the orca_mrci module. With ORCA 4.2, the FIC-CASPT2 approach is
available. The calling is in absolute analogy to the NEVPT2 branch and thus not
presented here. For more details see the manual.

Final notice: ORCA 4.0 has different frozen core settings compared to previous version.
For transition metal complexes, the 3s and 3p orbitals are now correlated. They are
quite important for the accurate energies and properties such as zero field splitting
(ZFS).5

1 C. Angeli, R. Cimiraglia, and J.-P. Malrieu, Chem. Phys. Lett. 350, 297 (2001).
2 C. Angeli, R. Cimiraglia, S. Evangelisti, T. Leininger, and J.-P. Malrieu, J. Chem. Phys. 114, 10252 (2001).
3 C. Angeli, R. Cimiraglia, and J.-P. Malrieu, J. Chem. Phys. 117, 9138 (2002).
4 Y. Guo, K. Sivalingam, E.F. Valeev, and F. Neese, J. Chem. Phys. 144, 94111 (2016).
5 K. Pierloot, Q.M. Phung, and A. Domingo, J. Chem. Theory Comput. 13, 537 (2017).
4

2 The [Cr(NH3)6]3+ complex – a standard run


Quit often we are interested in the ligand field spectrum of a transition metal complex.
As a simple example, we study a Cr(III) model complex. From elementary inorganic
chemistry, we know that Cr(III) is a d3-system and hence, the three metal d-electrons are
our primary focus. In later stage will explore various active spaces:

· CAS(3,5) consisting of the five 3d metal orbitals.


· CAS(7,7) includes two ligand orbitals.
· CAS(7,12) extended to include the second d-shell.

In this section, we demonstrate our standard approach to do such calculation. In the


appendix we find a slower paced description of the same example with much more
details on the selection of active orbitals/initial guess, state-averaging and basis-set
projection.

2.1 Organization of the Calculation: Preliminary Chemical Considerations

As mentioned above, the central Cr3+ ion is a d3 system. Since the [Cr(NH3)6]3+ complex is
very close to an octahedral coordination, we can expect the three metal t2g-based
orbitals to be singly occupied and hence a 4A2g ground state arises. Each NH3 provides a
single lone pair that can form a σ-bond with the central metal, but no π-bonds are
possible. Hence, we only need the two ligand orbitals of eg-symmetry when we want to
include the ligand orbitals in the active space.

For Cr, Ammonia acts as an intermediate-field ligand.6 If we are interested in the d-d
excited states of the system, the Tanabe-Sugano diagram (Figure 25) tell us, that we
expect 4T2g and two 4T1g excited states of the same multiplicity as the ground state. 7
Since T-terms are triply orbitally degenerate, 10 roots need to be calculated to capture
all spin allowed ligand field transitions. Of course, due to the presence of the hydrogens,
the actual symmetry is not strictly octahedral, but we use Oh group notation nonetheless.
According to the Tanabe-Sugano diagram, the next double states are derived from the
free ion 2G term, which splits into the triply degenerate 2T2, 2T1, the doubly degenerate
2E terms and an 2A term. Thus, we need a total of nine doublet roots.
1

6 L.G. Vanquickenborne, B. Coussens, D. Postelmans, A. Ceulemans, and K. Pierloot, Inorg. Chem. 30, 2978
(1991).
7 Derived from the 4F and 4P terms of the free ion.
5

Figure 1: Tanabe-Sugano diagram for d3 system in octahedral field

The initial geometry of the complex comes from a DFT geometry optimization, but it can
also be based on the crystal structure with just the hydrogens optimized.
# This is a slightly smoothed DFT optimized geometry in internal
# coordinates. This helps keeping the orbitals clean, as the ligands will
# be placed on the coordinates. Any xyz coordinates would do too, just
# replace “int” by “xyz” and give Cartesian Coordination (in Angström)
* int 3 4
Cr 0 0 0 0 0 0
N 1 0 0 2.137 0 0
N 1 2 0 2.137 90 0
N 1 2 3 2.137 90 90
N 1 2 3 2.137 90 180
N 1 2 3 2.137 180 180
N 1 2 3 2.137 90 270
H 2 1 3 1.041 114 0
H 2 1 3 1.041 114 120
H 2 1 3 1.041 114 240
H 3 1 2 1.041 114 0
H 3 1 2 1.041 114 120
H 3 1 2 1.041 114 240
H 4 1 2 1.041 114 315
H 4 1 2 1.041 114 195
H 4 1 2 1.041 114 75
H 5 1 2 1.041 114 0
H 5 1 2 1.041 114 120
H 5 1 2 1.041 114 240
H 6 1 4 1.041 114 270
H 6 1 4 1.041 114 30
H 6 1 4 1.041 114 150
H 7 1 2 1.041 114 45
H 7 1 2 1.041 114 165
H 7 1 2 1.041 114 285
*
It is advisable to align the molecule with respect to the coordinate axis e.g. in this
example the nitrogen atoms should be along the x,y,z-axis. This merely simplifies the
identification of the orbitals later on but has no influence on the mechanics of the
calculation. An exception is the ab-initio ligand field theory (AI-LFT),8 where the
alignment of the molecule in the axis frame is important.

2.2 Initial 3d CAS(3,5)


We start with the smallest active space that is just the 3d-metal orbitals. In later
stages, we can use these converged calculations to further extend the active space. In

8 Atanasov, M., Ganyushin, D., Sivalingam, K., Neese, F., Struct. and Bonding, (2012), 143:149-220
6

general, this is a good strategy for TM complexes with mostly ionic interactions
between metal and ligands, where the 3d-metal orbitals are strongly localized.

CASSCF calculations require the user to specify the number of active electron in active
orbitals. The program automatically starts with the “PModel” (~diagonalized LDA DFT
matrix) as initial guess. For many applications this is not a sufficient input as the active
orbitals are not consciously chosen. For transition metal complexes the desired 3d-metal
orbitals are often below HOMO-LUMO gap and hence do not automatically enter the
active space. The “PAtom” guess gives good atomic orbitals, with an extended Hückel
like ordering, which is a good idea for transition metal calculations as the desired metal
3d-orbitals are typically at the HOMO-LUMO gap and essentially of metal
character.

# Initial CASCF on [Cr(NH3)6]3+


#
# def2-TZVPP = triple zeta basis
# RI-JK = use the RI-JK approximation for the Fock matrix
# (not required – just used for more speed here)
# def2/JK = auxiliary basis for the RI approximation
# (automatically sets the auxiliary slots for
# the Fockian [AuxJK] and gradient integrals [AuxC])
# for Fock [AuxJK slot] and gradient integrals [AuxC slot])
#
# Conv = store integrals on disk (required for RIJK)
# XYZFile = leave coordinates on disk (convenient later)
# normalprint = slightly larger printing than default + includes Loewdin
# population analysis.
! def2-TZVPP def2/JK RI-JK Conv XYZFile PATOM
%casscf
nel 3 # number of active electrons
norb 5 # number of active orbitals
mult 4,2 # multiplicity blocks
nroots 10,9 # Roots per multiplicity blocks
end

* int 3 4
Cr 0 0 0 0 0 0
N 1 0 0 2.137 0 0
N 1 2 0 2.137 90 0
N 1 2 3 2.137 90 90
N 1 2 3 2.137 90 180
N 1 2 3 2.137 180 180
N 1 2 3 2.137 90 270
H 2 1 3 1.041 114 0
H 2 1 3 1.041 114 120
H 2 1 3 1.041 114 240
H 3 1 2 1.041 114 0
H 3 1 2 1.041 114 120
H 3 1 2 1.041 114 240
H 4 1 2 1.041 114 315
H 4 1 2 1.041 114 195
H 4 1 2 1.041 114 75
H 5 1 2 1.041 114 0
H 5 1 2 1.041 114 120
H 5 1 2 1.041 114 240
H 6 1 4 1.041 114 270
H 6 1 4 1.041 114 30
H 6 1 4 1.041 114 150
H 7 1 2 1.041 114 45
H 7 1 2 1.041 114 165
H 7 1 2 1.041 114 285
*

The above calculation the orbitals are state-averaged. In ORCA 4.0, the default state-
averaging sets equal weights for multiplicity blocks. The actual weights are also
printing in the output before the first CASSCF iteration, when the CI is setup.
7

In the following, the RI approximation is used to speed up the calculations. In Section


8.3, you will find more information on the accuracy of the RI approximation. The
keyword “xyzfile” produce a xyz coordinate file on disk that we use in later inputs
(cr_example1.xyz). The calculation converges in 8 iterations. The results are reported in
Table 2 with the extended active spaces.

2.3 Including ligand orbitals CAS(7,7)

In the second step, we improve the reference wave function by including the metal-
ligand bonding orbitals in the active space. Having the bonding and anti-bonding orbitals
should balance the active space.

The orbitals are sorted by their energies. Hence, the desired orbitals may not be the
highest doubly occupied orbitals. In fact, they are usually not. The highest ligand orbitals
are typically non-bonding, whereas we are looking for bonding orbitals that are
stabilized. Identifying these orbitals from the previous calculation requires looking at
the orbital coefficients or better to visualize the molecular orbitals. In this example,
the reduced Loewdin analysis printed at the end of the CAS(3,5) calculation is sufficient
to identify the ligand orbitals.9 The ligand orbitals of interest are the bonding partners of
the dz2 and dx2-y2 orbitals. In our example, we obtain the following canonical orbitals10:
34 35
-1.02677 -1.02677
2.00000 2.00000
-------- --------
0 Cr s 0.0 0.0
0 Cr dz2 0.0 25.2
0 Cr dx2y2 25.2 0.0
1 N s 0.7 0.2
1 N pz 0.0 0.0
1 N px 17.0 5.3
1 N py 0.0 0.0
2 N s 0.7 0.2
2 N pz 0.0 0.0
2 N px 0.0 0.0
2 N py 17.0 5.3
3 N s 0.0 0.9
3 N pz 0.0 22.3

Hence, we need to use the “rotate” feature in order to bring the correct orbitals on top
of the (formerly) doubly occupied space to include them in the active space of the next
calculation. The previously converged CAS(3,5) orbitals are denoted as “cas_5.gbw”.
Depending on the computed molecule, the bonding partner orbitals are not always
identified using the canonical orbitals. Since CASSCF is invariant to rotations within a
given subspace (internal, active or external), one might use a different representation,
where the identification process is simplified. The option “IntOrbs PMOs” leads to such a
representation and will be illustrated in the next version of the CASSCF tutorial.

Here is the ORCA input:

# Second step: include ligand orbitals

9 Identifying orbitals with the Loewdin analysis is bread and butter for CASSCF. A small parsing script that
filters all metal dominating orbitals might be handy.
10 The composition of the orbitals might change from calculation to calculation. In case of degenerate
orbitals each calculation returns an arbitrary mixture of the degenerate set.
8

#
! def2-TZVPP def2/JK RI-JK conv
! moread
%moinp "cas_5.gbw" # orbitals from the CAS(3,5) calculation

# The highest doubly occupied inactive orbitals are 37 and 38.


%scf
# swaps orbitals 34 and 35 to 37,38.
rotate {34,37} {35,38} end # longer notation: {34,37,90} {35,38,90}
end

%casscf nel 7
norb 7
mult 4,2
nroots 10,9
end

* xyzfile 3 4 cr_example1.xyz #previously generated xyzfile

This calculation converges in 5 iterations. The resulting gbw-file is denoted as


“cas_7.gbw”.

2.4 Including a second d-shell CAS(7,12)

In some cases, including a second d-shell can be useful to make the wave function more
flexible and obtain accurate results in conjunction with a subsequent second order
multireference perturbation method such as CASPT2 or NEVPT2.11 Most often it is not
necessary to include the entire second d-shell, but the ones that correspond to the
occupied 3d-metal orbitals.

To find the second d-shell, we use the keyword “extorbs doubleshell” in the converged
CAS(7,7) calculation. Based on the composition of the highest active orbital, the program
automatically identifies and produces a “second shell” in the vicinity of the active space.
It is important that the highest active orbital indeed has largest contribution from the
metal based d-orbital. For an active space consisting of 3d-metal orbitals the second
shell consists of the 4d-metal orbitals. For an active space consisting of 4d-metal orbital,
the second shell consists of 5d-metal orbitals and so on.

Note that the option does not work in conjunction with symmetry (UseSym). The
following input reads the converged CAS(7,7) orbitals and produces the second-d shell
(orbitals 44-48) in the correct order.
! def2-TZVPP def2/JK RI-JK conv
! moread
%moinp "cas_7.gbw" # orbitals from the cas(7,7) calculation
%casscf nel 7
norb 7
mult 4,2
nroots 10,9
extorbs doubleshell # produce the double-shell above the actives.
# all other virtuals are canonicalized
end
* xyzfile 3 4 cr_example1.xyz

As printed in the output, orbital nr.43 (highest active) is taken as reference and

11The second d-shell brings in a radial correlation effect that normally should be covered by the
dynamical correlation treatment. However, second order perturbation theory with a contracted first-
order interacting space is not flexible enough to provide this missing correlation. It is somewhat counter
the philosophy of the CASSCF method (or MCSCF in general) to include dynamic correlation in the active
space. However, it is common practice and hence described here.
9

the double-shell is produced in the MO range 44-48. Instead of the highest active orbital,
the active reference MO can be set with the additional keyword “DoubleShellMO 43”,
where 43 is the actual MO number.
---- THE CAS-SCF GRADIENT HAS CONVERGED ----
--- FINALIZING ORBITALS ---
---- DOING ONE FINAL ITERATION FOR PRINTING ----
--- Forming Natural Orbitals
--- Canonicalize Internal Space
--- SortedExt: Largest compononent of the highest active orbital (Nr. 43) on atom 0 Cr with l=2
--- SortedExt: Double Shell Range 44 -> 48

We confirm the correctness by inspecting the Loewdin population analysis and


visualizing the orbitals (see the next subsections). The associated gbw file is denoted as
“cas_7_sorted.gbw” in the next step.
43 44 45 46 47
-0.57548 0.26599 0.26614 0.26629 0.77121
0.60166 0.00000 0.00000 0.00000 0.00000
-------- -------- -------- -------- --------
0 Cr dxz 49.4 0.0 45.2 45.2 0.0
0 Cr dyz 49.4 0.0 45.2 45.2 0.0
0 Cr dx2y2 0.0 0.0 0.0 0.0 69.3
0 Cr dxy 0.0 90.4 0.0 0.0 0.0

Having generated a double-shell, we will setup the calculation for the extended active
space. Since we start from an already converged CASSCF wave function, we may try the
Newton-Raphson method (keyword “switchstep nr”) to obtain convergence here. The
rate of convergence is higher with this method, but the radius of convergence is smaller.
The program can use two different convergers specified with “orbstep” and
“switchstep”. Far off from convergence “orbstep” is used. The SuperCI is good choice for
large initial gradients. ORCA changes the converger to “Switchstep” when the calculation
is close to convergence (||g|| < 0.02).12 The NR method is a safe pick for re-converging
calculations that have already been converged with a slightly different active space or
basis set.
! def2-TZVPP def2/JK RI-JK conv
! moread
%moinp "cas7_sorted.gbw" # cas(7,7) orbitals with prepared virtual space.

%casscf nel 7
norb 12 #3d + ligands + 4d orbitals
mult 4,2
nroots 10,9
cistep accci # faster, more memory hungry algorithm for the CI step
switchstep nr
end

* xyzfile 3 4 cr_example1.xyz

In many cases, switching to the computationally more demanding NR solver does not
result in net time savings. In this example, the “switchstep NR” and the default converger
perform equally well (4-6 iterations).

For larger active spaces or many roots, the timings can be considerably improved using
the “CIStep ACCCI” for the CI calculation. The method is absolutely equivalent to the
default CI solver, but uses are more memory demanding algorithm. The final set of
orbitals is denoted as “cas_12.gbw” in the next section.

12 controlled by the keyword “switchconv”


10

The orbitals and the occupation numbers from the converged calculation are shown
below. They are ordered by increasing occupation number. You see an ideal shape and
ordering of the orbitals, which makes the interpretation of the results very convenient. It
is also a quality control for your calculation to ensure that you have arrived at the
desired enlarged active space. Note that not all of the orbitals are perfectly aligned
to the coordinating ligands. This is perfectly normal as some of the orbitals are
degenerate and hence arbitrarily mixed.

Figure 2: These two orbitals are the antibonding eg-counterparts in the second d-shell. Notice how large these
orbitals are. If we would plot a radial cut, you would observe that they have a node, whereas the primary d-
orbitals do not have it. (isosurface value 0.05)

Figure 3: These three orbitals are the second d-shell counterparts of the nonbonding t 2g based metal orbitals.

Figure 4: These three orbitals are the nonbonding metal d-orbitals of t 2g origin.
11

Figure 5: These are the antibonding eg based orbitals in the primary metal d-based set.

Figure 6: These two orbitals are the essentially doubly occupied bonding counterpart of the metal eg-orbitals

2.5 Reading the wave function


The actual weights are also printing in the output before the first CASSCF iteration, when
the CI is setup. Let us look at some of the results of this calculation. After convergence,
you find the definition of the wave function for each state:
---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 4 NROOTS=10
---------------------------------------------
ROOT 0: E= -1379.8845743311 Eh
0.96606 [ 0]: 221110000000
0.00466 [ 2552]: 111111100000
0.00444 [ 165]: 211111000000
0.00444 [ 1828]: 121110100000
0.00276 [ 809]: 201112000000
0.00276 [ 6660]: 021110200000
ROOT 1: E= -1379.7943027489 Eh 2.456 eV
0.73212 [ 1]: 221101000000
0.24403 [ 9]: 221010100000
0.00277 [ 1835]: 121101100000
0.00257 [ 6687]: 021101200000
...

The program then lists the main contributing configurations that are active space
occupation patterns. For example, the ground state has a weight of 0.96606, which
means that the lowest root is dominated to 96.6% by a single configuration (this number
is the sum of the squares of the CI coefficients for all configuration state functions that
belong to this configuration, e.g. the linearly independent spin couplings). This
configuration has the active space occupation pattern 221110000000 which means the
first active orbital is doubly occupied, the second doubly occupied as well, the next three
12

orbitals are singly occupied and the remaining orbitals are empty. If you look at your
orbitals (see above), you see that the first two are the ligand based bonding orbitals, the
next three the metal t2g based orbitals, followed by the two metal eg based orbitals and
the remaining ones are the second d-shell.13 ORCA by default uses natural orbitals for
the active space. The metal eg based orbitals have a slightly higher occupancy due the
presence of the ligand orbitals.
Repeating the calculation, you should be aware of degeneracy, which can lead to
different outputs (orbitals and wavefunction composition).

Note that ORCA employs configuration state functions. Occasionally one interested in
the CI Coefficients or the representation in terms of spin determinants. This is
possible with the keyword “PrintWF” and discussed in Section 8.5 in more detail.

The program then prints the CASSCF transition energies:


-----------------------------
SA-CASSCF TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0 ,MULT 4) = -1379.884565085 Eh -37548.568 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 0 2 0.086678 2.359 19023.6
2: 1 2 0.086685 2.359 19025.2
3: 2 2 0.090122 2.452 19779.4
4: 3 2 0.090126 2.452 19780.4
5: 4 2 0.090128 2.452 19780.7
6: 1 4 0.090272 2.456 19812.3
7: 2 4 0.090281 2.457 19814.3
8: 3 4 0.090297 2.457 19817.9
9: 5 2 0.125763 3.422 27601.8
10: 6 2 0.125771 3.422 27603.5
11: 7 2 0.125778 3.423 27605.1
12: 4 4 0.128989 3.510 28309.8
13: 5 4 0.128990 3.510 28310.1
14: 6 4 0.129020 3.511 28316.7
15: 8 2 0.160774 4.375 35285.7
16: 7 4 0.202971 5.523 44547.1
17: 8 4 0.202979 5.523 44548.6
18: 9 4 0.203002 5.524 44553.7

2.6 NEVPT2 CAS(7,12)


CASSCF calculations typically do not accurately reproduce excitation energies. The
easiest way to improve the results is with NEVPT2. It requires one additional keyword
on top the CASSCF input. In this example, we employed the SC-NEVPT2 method using
the RI approximation. For larger molecules (>80 atoms), we recommend the DLPNO-
NEVPT2 approach, which is linear scaling extension to the FIC-NEVPT2 method.14 The
accuracy of the RI approximation as well as DLPNO calculations are further reported in
Section 8.3 of this tutorial.15
! def2-TZVPP def2/JK RI-JK conv RI-NEVPT2
! moread
%moinp "cas12.gbw" # converged cas(7,12) orbitals.

%casscf nel 7
norb 12 #3d + ligands + 4d orbitals

13 The number in square brackets is the number of the configuration in the configuration list and is
irrelevant.
14 Y. Guo, K. Sivalingam, E.F. Valeev, and F. Neese, J. Chem. Phys. 144, 94111 (2016).
15 With ORCA 4.1 auxiliary basis sets are assigned to slots. Here, !RIJK sets the def2/JK auxiliary basis to
the AuxJK (Fock construction) and AuxC (orbital gradient, NEVPT2) slots.
13

mult 4,2
nroots 10,9
end

* xyzfile 3 4 cr_example1.xyz

There is a fair amount of output generated in the course of the calculation that, most of
the time, is of limited interested to the user. However, eventually we reach the section:
===============================================================
NEVPT2 Results
===============================================================

For the really curious user the program then prints the contributions of each excitation
class to the final NEVPT2 correction. Finally, we obtain:
-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------
LOWEST ROOT (ROOT 0, MULT 4) = -1381.655466666 Eh -37596.757 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 1 2 0.075680 2.059 16609.9 #2Eg
2: 0 2 0.075724 2.061 16619.6
3: 2 2 0.078512 2.136 17231.4 #2T1g
4: 3 2 0.078513 2.136 17231.5
5: 4 2 0.078542 2.137 17238.0
6: 3 4 0.096108 2.615 21093.2 #4T2g
7: 1 4 0.096168 2.617 21106.3
8: 2 4 0.096178 2.617 21108.6
9: 7 2 0.113651 3.093 24943.6 #2T2g
10: 5 2 0.113654 3.093 24944.3
11: 6 2 0.113660 3.093 24945.4
12: 4 4 0.128469 3.496 28195.7 #4T1g
13: 5 4 0.128472 3.496 28196.3
14: 6 4 0.128505 3.497 28203.6
15: 8 2 0.157155 4.276 34491.6 #2A1g
16: 7 4 0.207218 5.639 45479.2 #4T1g
17: 8 4 0.207328 5.642 45503.2
18: 9 4 0.207356 5.642 45509.4
These are the final transition energies. We see that the lowest state is a quartet (as
expected) and the next higher states are doublets starting at around 16600 cm-1. Hence,
we have an energetically well-isolated ground state. The states can be readily assigned
with the aid of their degeneracy and the Tanabe-Sugano diagram for d3 as indicated
above.

For convenience, the results of the aforementioned calculations are summarized Table 1.
It is evident that the changes from CASSCF to NEVPT2 are not enormous and amount
approximately 0.3 eV. This indicates that the CASSCF description of the spectrum is
already pretty good and the NEVPT2 results are reliable. Another interpretation can be
that static electron correlation is dominating in this Cr complex; hence the recovered
dynamic electron correlation doesn’t change the result much. Indeed, as far as
comparison to experiment is possible, the results are within 0.23 eV. This is a good
result given that the relativistic effect was neglected, environment effects not included
and no attempt has been made to reach the basis set limit. Typically, the low-spin states
have the largest errors and profit the most from larger basis sets and any improvement
in the treatment of the dynamical correlation.

Table 1. Few energy ligand field spectra using the default weighting (equal weights for multiplicity blocks).
The CAS(7,12) consists of the 3d orbitals, 2 ligand orbitals and the second d-shell.

State Exp.a CASSCF NEVPT2


14

(7,12) (7,12)
2E 15300 19023 16609
g
2T 15300 19779 17231
1g
2T - 27601 24943
2g
4T 21550 19812 21093
2g
4T 28500 28309 28195
1g
2A - 35285 34491
1g
4T - 44547 45479
1g
a Jorgensen,
C., K., Absorption Spectra and Chemical Bonding in Complexes, Pergamon Press, Oxford, 1962,
p291 and references therein.

Let us investigate the influence of state-averaging and the extension of the active space
on the ligand field spectrum computed with NEVPT2. For comparison, we provide the
following results in Table 2:
· Results with the minimal SA-CAS(3,5) that is just 3d-metal orbitals. The orbitals
are optimized for the quartet states
· SA-CAS(3,5) with the default weighting for the orbital optimization.
· Results with two ligand orbitals included in the active space: SA-CAS(7,7)
· Results with the complete second d-shell included: SA-CAS(7,12).

Averaging over just the quartets or the quartets and the doublets has a minor effect
(<0.05 eV), which is expected as the optimal orbitals for all ligand field states are
expected to be fairly similar. In this example, inclusion of the ligand orbitals barely
changes the spectra. The effect of ligand orbitals is more important for covalent ligands.
The effect of the second d-shell is recognizable but not overwhelmingly large. The
second d-shell is more pronounced for complexes, where the 3d orbitals are strongly
occupied.

Table 2: NEVPT2 Vertical Transition energies of [Cr(NH3)6]2+ in cm-1. The calculation involves 10 quartet and
9 doublet states. CAS(7,7) includes ligand orbitals. CAS(7,12) includes ligand orbitals and the second d-shell.

State NEPT2 NEVPT2 NEVPT2 NEVPT2 Exp.c


CAS(3,5)a CAS(3,5)b CAS(7,7)b CAS(7,12)b
2E 17117 17056 17191 16606 15300
g
2T 18001 17926 17873 17228 15300
1g
2T 25965 25878 25757 24939 -
2g
4T 20324 20558 21000 21095 21550
2g
4T 28828 28999 28379 28198 28500
1g
2A 34369 34533 34968 34492 -
1g
4T 43959 44406 45209 45483 -
1g
a Orbitals optimized for the quartet states.
b Orbitals optimized with equal weights for quartets and doublets.
c Jorgensen, C., K., Absorption Spectra and Chemical Bonding in Complexes, Pergamon Press, Oxford, 1962,
p291 and references therein.

3 [Cr(NH3)6]3+ complex - extracting ligand field parameters.


15

The ab initio ligand theory (AILFT) developed at the MPI-CEC allows one to deduce the
ligand field parameters for a given complex directly from the ab initio calculation.16,17
For the nearly octahedral complex here, this boils down to the ligand-field splitting
10Dq and the Racah parameters B and C. The Racah parameters B and C, used to
describe the inter-electronic repulsion within the d-shell, are related with the Slater-
Condon parameters as follows:
B < F2 , 5F4 ,
C < 35·F4 ,
where F2 and F4 are the normalized Slater-Condon parameters18 with the following
normalization factors
F2 < F 2 / 49,
F4 < F 4 / 441.

The connection between ab initio and ligand field theory is very strong and provides a
lot of insight. The ligand field parameters that come out of the calculation are chosen
such that the ligand field full-CI treatment provides results as close to the ab initio
results as possible within the limitations of the ligand field model itself. Hence, though
the reproduction is not perfect, within the domain of applicability of ligand field theory,
it is usually very good (excited state energies are typically reproduced with an accuracy
of a few hundred wavenumbers). The origin of the limitations of ligand field theory is
mainly the anisotropic covalency of the metal-ligand bond.19 This provides anisotropy in
the electron-electron repulsion, which is not reflected in the Racah parameters. Hence,
in this section we compare how dynamical correlation in form of the NEVPT2 correction
influences the results for the ligand field parameters.

Ab initio ligand field results are computed and printed out by the keyword “actorbs
dorbs” in the CASSCF block. In the spirit of the ligand field theory, the active space is
restricted to the 3d-metal orbitals. Furthermore, all roots of a given multiplicity must
be included. If NEVPT2 is requested, the ligand field parameters for both methods are
listed in the output.
# Extract the LFT parameters
! def2-TZVPP def2/JK RI-JK conv PATOM
%casscf nel 3
norb 5
mult 4,2 # quartet and doublet multiplicities
nroots 10,40 # 10 quartets, 40 doublets
# you can adjust the weight of each manually
# as described in the manual. Default is equal
# weights.

PTMethod SC_NEVPT2 # invoke the SC-NEVPT2 correction


actorbs dorbs # invokes the ab initio LFT analysis.
# CAS must be 3d orbitals!
end

* xyzfile 3 4 cr_example1.xyz

16 Atanasov, M., Ganyushin, D., Sivalingam, K., Neese, F., Struct. and Bonding, (2012), 143:149-220
17 Atanasov,M.;Zadrozny,J.M.; Jeffrey R. Long, J.R.; Neese, F., Chem.Sci. (2013), 4,139-156.
18 Notation: subscript = normalized, superscript = unormalized
19 S.K. Singh, J. Eng, M. Atanasov, and F. Neese, Coordination Chemistry Reviews 344, 2 (2017).
16

After the usual CASSCF and NEVPT2 printings, the output contains the AILFT analysis,
where the reference is given in brackets e.g. for the CASSCF spectrum the ligand field
matrix takes the following form.
------------------------------
AILFT MATRIX ELEMENTS (CASSCF)
------------------------------

E0 = 0.057992449 a.u. = 1.578 eV = 12727.9 cm**-1


H(dxy ,dxy )= 0.004776019 a.u. = 0.130 eV = 1048.2 cm**-1
H(dyz ,dxy )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(dyz ,dyz )= 0.004804947 a.u. = 0.131 eV = 1054.6 cm**-1
H(dz2 ,dxy )= 0.000002581 a.u. = 0.000 eV = 0.6 cm**-1
H(dz2 ,dyz )= 0.000000001 a.u. = 0.000 eV = 0.0 cm**-1
H(dz2 ,dz2 )= 0.079801002 a.u. = 2.171 eV = 17514.3 cm**-1
H(dxz ,dxy )= -0.000000001 a.u. = -0.000 eV = -0.0 cm**-1
H(dxz ,dyz )= 0.000007713 a.u. = 0.000 eV = 1.7 cm**-1
H(dxz ,dz2 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(dxz ,dxz )= 0.004804937 a.u. = 0.131 eV = 1054.6 cm**-1
H(dx2-y2,dxy )= 0.000000003 a.u. = 0.000 eV = 0.0 cm**-1
H(dx2-y2,dyz )= 0.000000050 a.u. = 0.000 eV = 0.0 cm**-1
H(dx2-y2,dz2 )= 0.000000005 a.u. = 0.000 eV = 0.0 cm**-1
H(dx2-y2,dxz )= -0.000000052 a.u. = -0.000 eV = -0.0 cm**-1
H(dx2-y2,dx2-y2)= 0.079790442 a.u. = 2.171 eV = 17512.0 cm**-1
B = 0.004650608 a.u. = 0.127 eV = 1020.7 cm**-1
C = 0.017752382 a.u. = 0.483 eV = 3896.2 cm**-1 (C/B= 3.82)

Diagonalization of the ligand field matrix defines the LFT orbitals. Unlike CASSCF natural
orbitals, the LFT orbitals are perfect 3d orbitals that rest in the eigenframe of the
molecule. The energy splitting of the d-d excited states translates into differences in
orbital energies.
The ligand field one electron eigenfunctions:
Orbital Energy (eV) Energy(cm-1) dxy dyz dz2 dxz dx2-y2
1 0.000 0.0 -1.000000 -0.000001 0.000034 -0.000033 0.000000
2 0.001 4.7 0.000022 0.706880 -0.000000 -0.707334 -0.000001
3 0.001 8.0 0.000024 -0.707334 0.000000 -0.706880 -0.000000
4 2.041 16463.8 0.000000 0.000001 -0.000485 -0.000001 1.000000
5 2.042 16466.1 -0.000034 -0.000000 -1.000000 0.000000 -0.

Ligand field orbitals were stored in cram5.casscf.lft.gbw

As highlighted in the snippet above, these orbitals are stored in the conventional .gbw
format. Like any other set of orbitals, they can be visualized or analyzed by means of
population analysis.

The AILFT module re-computes the energies of the states using the extracted LF
parameters. The ligand field spectrum is compared with the ab initio results abbreviated
with AI.
------------------------------------------------
COMPARISON OF AB INITIO AND LIGAND FIELD RESULTS
------------------------------------------------

Block 1
---------
AI-Root 0: E(AI)= 0.000 eV -> LF-Root 0: 0.000 eV S= 1.000 Delta= 0.000 eV
AI-Root 1: E(AI)= 2.096 eV -> LF-Root 1: 2.041 eV S= 1.000 Delta= 0.056 eV
AI-Root 2: E(AI)= 2.096 eV -> LF-Root 2: 2.041 eV S= 1.000 Delta= 0.056 eV
AI-Root 3: E(AI)= 2.097 eV -> LF-Root 3: 2.041 eV S= 1.000 Delta= 0.056 eV
AI-Root 4: E(AI)= 3.193 eV -> LF-Root 4: 3.127 eV S= 1.000 Delta= 0.066 eV
AI-Root 5: E(AI)= 3.193 eV -> LF-Root 5: 3.127 eV S= 1.000 Delta= 0.065 eV
AI-Root 6: E(AI)= 3.193 eV -> LF-Root 6: 3.128 eV S= 1.000 Delta= 0.066 eV
AI-Root 7: E(AI)= 5.021 eV -> LF-Root 7: 4.893 eV S= 1.000 Delta= 0.128 eV
AI-Root 8: E(AI)= 5.021 eV -> LF-Root 8: 4.894 eV S= 1.000 Delta= 0.128 eV
AI-Root 9: E(AI)= 5.022 eV -> LF-Root 9: 4.894 eV S= 1.000 Delta= 0.128 eV
RMS error for this block = 0.084 eV = 681.0 cm**-1Block 1
17

In the snippet above, the average deviation between LFT and ab initio results is in the
order of 0.084 eV. This beautifully demonstrates the agreement between LFT and
CASSCF.

Completing the output for CASSCF, the analysis is repeated for the NEVPT2 results
starting with the header below.
------------------------------
AILFT MATRIX ELEMENTS (NEVPT2)
------------------------------

E0 = 0.062633346 a.u. = 1.704 eV = 13746.4 cm**-1


H(dxy ,dxy )= 0.005516301 a.u. = 0.150 eV = 1210.7 cm**-1
H(dyz ,dxy )= 0.000000003 a.u. = 0.000 eV = 0.0 cm**-1
H(dyz ,dyz )= 0.005531474 a.u. = 0.151 eV = 1214.0 cm**-1
H(dz2 ,dxy )= 0.000003987 a.u. = 0.000 eV = 0.9 cm**-1
H(dz2 ,dyz )= -0.000000001 a.u. = -0.000 eV = -0.0 cm**-1
H(dz2 ,dz2 )= 0.085674068 a.u. = 2.331 eV = 18803.3 cm**-1
H(dxz ,dxy )= -0.000000003 a.u. = -0.000 eV = -0.0 cm**-1
H(dxz ,dyz )= 0.000007317 a.u. = 0.000 eV = 1.6 cm**-1
H(dxz ,dz2 )= 0.000000001 a.u. = 0.000 eV = 0.0 cm**-1
H(dxz ,dxz )= 0.005531464 a.u. = 0.151 eV = 1214.0 cm**-1
H(dx2-y2,dxy )= 0.000000003 a.u. = 0.000 eV = 0.0 cm**-1
H(dx2-y2,dyz )= 0.000000056 a.u. = 0.000 eV = 0.0 cm**-1
H(dx2-y2,dz2 )= 0.000000007 a.u. = 0.000 eV = 0.0 cm**-1
H(dx2-y2,dxz )= -0.000000054 a.u. = -0.000 eV = -0.0 cm**-1
H(dx2-y2,dx2-y2)= 0.085646730 a.u. = 2.331 eV = 18797.3 cm**-1
B = 0.004529762 a.u. = 0.123 eV = 994.2 cm**-1
C = 0.014136482 a.u. = 0.385 eV = 3102.6 cm**-1 (C/B= 3.12)

Dynamical correlation changes C, while B is remains almost unchanged. As expected the


deviation of the LFT spectrum from the NEVPT2 results is larger compared to the
CASSCF findings reported earlier.
------------------------------------------------
COMPARISON OF AB INITIO AND LIGAND FIELD RESULTS
------------------------------------------------

Block 1
---------
AI-Root 0: E(AI)= 0.000 eV -> LF-Root 0: 0.000 eV S= 1.000 Delta= 0.000
eV
AI-Root 1: E(AI)= 2.566 eV -> LF-Root 1: 2.180 eV S= 1.000 Delta= 0.385 eV
AI-Root 2: E(AI)= 2.572 eV -> LF-Root 2: 2.180 eV S= 1.000 Delta= 0.392 eV
AI-Root 3: E(AI)= 2.573 eV -> LF-Root 3: 2.181 eV S= 1.000 Delta= 0.392 eV
AI-Root 4: E(AI)= 3.620 eV -> LF-Root 4: 3.282 eV S= 1.000 Delta= 0.338 eV
AI-Root 5: E(AI)= 3.620 eV -> LF-Root 5: 3.282 eV S= 1.000 Delta= 0.338 eV
AI-Root 6: E(AI)= 3.627 eV -> LF-Root 6: 3.282 eV S= 1.000 Delta= 0.344 eV
AI-Root 7: E(AI)= 5.530 eV -> LF-Root 8: 5.108 eV S= 1.000 Delta= 0.421 eV
AI-Root 8: E(AI)= 5.535 eV -> LF-Root 7: 5.108 eV S= 1.000 Delta= 0.427 eV
AI-Root 9: E(AI)= 5.536 eV -> LF-Root 9: 5.108 eV S= 1.000 Delta= 0.427 eV
(Dropping RMS error for the reference energy)
RMS error for this block = 0.387 eV = 3117.9 cm**-1

We note that the AILFT module can extract the spin-orbit coupling parameter z, when
the spin-orbit coupling (SOC) correction is requested in the CASSCF block.
# spin-orbit coupling corrected spectrum and extraction of “Zeta”
! def2-TZVPP def2/JK RI-JK conv

%casscf nel 3
norb 5
mult 4,2 # quartet and doublet multiplicities
nroots 10,40 # 10 quartets, 40 doublets
# you can adjust the weight of each manually
# as described in the manual. Default is equal
# weights.
PTMethod SC_NEVPT2 # invoke the SC-NEVPT2 correction
actorbs dorbs # invokes the ab initio LFT analysis.
18

# CAS must be 3d orbitals!


rel
dosoc true # included SOC using QDPT
end
end

* xyzfile 3 4 cr_example1.xyz

As described in the manual in more detail, the input above produces SOC corrected
spectrum and zero-field splitting parameters for CASSCF and NEVPT2. At the end of the
AILFT output section, the program prints z parameters derived from a fit to the CASSCF
SOC integrals.
----------------------------------------------
SPIN ORBIT COUPLING (based on CASSCF orbitals)
----------------------------------------------
# printing of the ab initio soc integrals omitted here
...
Fit to the SOC matrix elements
a = 15.000000
b = 0.482 eV = 3885.3 cm**-1
SOC constant zeta = 0.032 eV = 259.0 cm**-1
# printing of the lft based soc integrals omitted here
...
RMS error of nonzero matrix elements = 2.5 cm**-1

As reflected by the root mean square error (RMS) the consistency between the ab initio
SOC integrals and the parameterized SOC integrals is impressive!

4 [CrCl6]3- model complex - CASSCF for larger active spaces


The electronic absorption spectrum of octahedral CrCl63- recorded in Cr3+ doped
Cs2NaScCl6 (Figure 7 ) displays all three spin allowed d-d transitions, and in addition, the
transitions from the 4A2 ground state into the 2Eg, 2T1g and 2T2g all due to spin flip
transitions within the t2g3 ground state configuration.
19

Figure 7. 15 K absorption spectrum of 4.1% Cr(III): Cs2NaScCl6.

The design of the calculation follows similar considerations as in the hexamine complex
studied in the previous section. However, the chlorine ligand has -orbitals available for
bonding and also forms a more covalent ligand bond. Hence, we include the -type
ligand orbitals. Due to the larger active space, the calculation is computationally more
demanding than the amine system studied earlier. It is a good example to discuss a few
option designed for larger active spaces.

The complex has a large negative charge and thus a gas-phase calculation is certainly not
the best choice. A good computational protocol should take into account the
environment effect for example by considering an ECP embedding together with point
charges and much larger basis set. 20 For now we proceed with the gas-phase calculation.

Following the protocol described in Section 2, the molecule is carefully placed into the
xyz axis frame. The geometry is stored as “crcl6-03.xyz”. The CAS(3,5) calculation state-
averaged over 10 quartet roots and 9 doublet roots converges smoothly with the PAtom
guess.
! SV def2/JK RI-JK conv PAtom xyzfile
%MaxCore 1000
%casscf nel 3
norb 5
mult 4,2
nroots 10,9
end
*xyz -3 4
Cr 0.000000 0.000000 0.000000
Cl 2.467400 0.000000 0.000000

20D. Maganas, M. Roemelt, M. Hävecker, A. Trunschke, A. Knop-Gericke, R. Schlögl, and F. Neese, Phys.
Chem. Chem. Phys. 15, 7260 (2013).
20

Cl 0.000000 2.467400 0.000000


Cl 0.000000 0.000000 -2.467400
Cl 0.000000 -2.467400 -0.000000
Cl -2.467400 0.000000 0.000000
Cl 0.000000 -0.000000 2.467400
*

Thereafter, we visually inspect the doubly occupied space and identify the and -
bonding ligand orbitals depicted in Figure 8.

Figure 8. Ligand orbitals selected from the converged CAS(3,5) calculation. In our output these are the
orbitals 46, 47, 51, 52 and 53.

For the CASSCF orbitals, there is a little mixing between -ligand orbitals and the 3d-
metal orbitals. Their inclusion in the active space most probably does not affect the d-d
spectrum. Nevertheless, we include the full set of ligand orbitals and run the CAS(13,10)
calculation. The keyword “extorb doubleshell” is set in preparation of the next step that
is the inclusion of the second d-shell.
! SV def2/JK RI-JK conv moread
%moinp “cas_5.gbw” # converged CAS(3,5) orbitals

%MaxCore 1000
# rotated ligand orbitals to be included in the active space 58-67
%scf rotate {47,61}{46,62}{53,58}{52,59}{51,60} end end
%casscf nel 13
norb 10
mult 4,2
nroots 10,9

extorbs doubleshell # produces the second d-shell


cistep accci # faster for multi-root calculations
end
*xyzfile -3 4 crcl6-03.xyz

We further extend the active space by the second d-shell orbitals. For the input above,
these are the next five virtual orbitals. Inclusion of the entire second d-shell leads to a
CAS(13,15), which is on the verge of the doable, but is a demanding calculation. To
reduce the computational complexity, one could restrict the double-shell effect on the
stronger occupied 3d-metal orbitals (t2g according to natural orbitals). In this case, the
smaller CAS(13,13) needs to be considered. For CASSCF calculations with larger active
spaces, ORCA features two spin-adapted alternatives to conventional full CI solvers
21

· Density Matrix Renormalization Group approach (DMRG) developed in the Chan


group.21
· Iterative Configuration Expansion (ICE), which is a variant of the CIPSI method
proposed by Malrieu and coworkers.

Both approaches are approximate solutions to the full-CI problem. Hence, less tight
CASSCF convergence thresholds are sufficient and in fact recommended (etol 1e-
6). The setup of DMRG typically requires a bit more insight. There are excellent reviews
on the subject.22,23 Here, we focus on the ICE approach. Note that the accuracy of ICE and
DMRG can be systematically controlled – see the manual for more details. We calibrate
the accuracy of the ICE methodology using the default settings for the CAS(13,13) before
applying it to the larger CAS(13,15).
! SV def2/JK RI-JK conv
! moread
%maxcore 4000
%moinp "cas_10.gbw" # converged cas(13,10) calculation with the
# double shell ranging from 68-72 (t2g orbitals first)
%casscf nel 13
norb 13 #or 15 if the full second d-shell is included
mult 4,2
nroots 10,9
etol 1e-6 # default = 1e-7
cistep ice # approximate CI step for large active spaces
end
*xyzfile -3 4 crcl6_03.xyz

The results are summarized Table 3 together with the experimental values.24 The
ICE(13,13) and the exact CAS(13,13) are practically identical. The extended ICE(13,15)
improves the results in particular for quartet excitations. Already at this level of theory,
the quartet transitions are in good agreement. The doublet excitations (low-spin)
should be further corrected with inclusion of more dynamic electron correlation.
To compare with experimental results, we should enlarge the basis and treat
environment effects.

Table 3. d-d transition energies for the [CrCL6]3- model complex. All calculations are done in the SV basis.

Excitations Exp. CAS(13,13) ICE(13,13) ICE(13,15)


4T 11900 13954 13954 12477
2
2E 13545 19100 19099 19059
2T 14180 19700 19698 19659
1
4T (1) 18500 20110 20104 18442
1
2T 20250 25923 25922 25452
2
2A - 29964 29963 28354
1
4T (2) 28500 31352 31348 28520
1

21 Sharma, S.; Chan, G. K.-L. (2012) J. Chem. Phys., 136, 124121


22 T. Yanai, Y. Kurashige, W. Mizukami, J. Chalupský, T.N. Lan, and M. Saitow, Int. J. Quantum Chem. 115,
283 (2015).
23 R. Olivares-Amaya, W. Hu, N. Nakatani, S. Sharma, J. Yang, and G.K.-L. Chan, J. Chem. Phys . 142, 34102
(2015).
24 O.S. Wenger, H.U. Güdel, J. Chem. Phys. 114, 5832-5841, 2001
22

5 A comment on using ANO basis sets


Many practitioners of CASSCF are used to employ ANO basis sets. In fact, ANOs have a
lot to recommend themselves. They are accurate and systematically extendable and they
form an excellent basis for correlated calculations. The drawback of ANOs is the
intrinsically high computational cost due to the large number of primitives in which the
orbitals are expanded. While ORCA is certainly not fully optimized for ANO calculations,
there is an integral program (orca_anoint) that makes use of the general contraction
scheme that underlies ANO construction. There are a few tricks to help those
calculations.

First let us recall the ANO basis sets that are built into ORCA:
# The ORCA ANO basis sets

# These are our own ANO basis sets described in


# Neese, F.; Valeev, E. F. Revisiting the Atomic Natural Orbital Approach
# for Basis Sets: Robust Systematic Basis Sets for Explicitly Correlated
# and Conventional Correlated ab initio Methods. J. Chem. Theory Comput.
# 2011, 7, 33-43.
# n= D, T, Q, 5, 6
! ano-pVnZ
# similar, slightly extended ANO sets
! aug-pVnZ
# augmented with one more shell except for the highlest L-quantum number
! saug-pVnZ
# augmented with more s-functions

# Other useful built-in ANO basis sets.


# A complete list is reported in the manual section 9.3.1
! ANO-RCC-DZP
! ANO-RCC-TZP
! ANO-RCC-FULL

If the desired ANO basis is not available in ORCA, you can read or define the basis in the
%basis block. Reading a basis from the EMSL is straight forward. Just select the
elements and the “GAMESS US” format. Then copy and paste the basis set information in
a text-file. In that case, it is very important to set the flag “ANOBasis true”!

# Reading your own ANO basis set e.g. from the EMSL
%basis GTOName “MyANOFile.bas”
# for the format check the manual!
# it is essentially EMSL Gamess US format
ANOBasis true # this tells the program that it deals with ANOs
# !! IMPORTANT – YOU CAN NOT MIX ANO AND NON-ANO BASES!!

Now, let us revisit the [CrCL6]3- example from Section 4 with ANO basis sets. We use the
resolution of the identity approximation and conventional integral storage in order
to speed up the calculation. For small basis sets this should pretty much always be
possible, even if the molecules are big. The PAtom guess is not available for ANO basis
sets. Hence, we start with the default guess, inspect the orbitals and rotate accordingly
(guess.gbw file).
# ANO calculation
#
# ANO-RCC-DZP : double zeta ANO-RCC basis (should be used with DKH)
# RI-JK : Use fitting for all integrals (critical for performance)
# Conv : Store integrals on disk (critical for performance)

!ANO-RCC-DZP DKH AutoAux ri-jk conv moread


%moinp “guess.gbw” # rotated PModel guess with 3d orbitals active
%casscf nel 3
23

norb 5
nroots 10
end
* int -3 4
Cr 0 0 0 0.0000 0.000 0.000
Cl 1 0 0 2.4674 0.000 0.000
Cl 1 2 0 2.4674 90.000 0.000
Cl 1 2 3 2.4674 90.000 90.000
Cl 1 2 3 2.4674 90.000 180.000
Cl 1 2 3 2.4674 180.000 0.000
Cl 1 2 3 2.4674 90.000 270.000
*

There are no pre-defined auxiliary basis set for ANO-RCC. Thus we use the AutoAux
construction, which generates a big decontracted auxiliary basis set that can be used in
CASSCF / NEVPT2 calculations.25 From there on, everything is pretty much the same as
in Section 4. However, at the DZP level, including the ligand orbitals we observe a
trailing convergence as can be seen from the gradient progression below.
||g|| = 0.103578893 Max(G)= 0.031317337 Rot=134,38
||g|| = 0.027803852 Max(G)= 0.005852422 Rot=132,8
||g|| = 0.010459608 Max(G)= 0.002120086 Rot=141,38
... 20 iterations more ...
||g|| = 0.044716665 Max(G)= -0.026700552 Rot=65,35
||g|| = 0.022609157 Max(G)= -0.015214902 Rot=65,35
||g|| = 0.019453693 Max(G)= -0.011320410 Rot=65,35
||g|| = 0.028508328 Max(G)= 0.019778929 Rot=65,35
||g|| = 0.014822140 Max(G)= -0.006988486 Rot=61,35
||g|| = 0.018136280 Max(G)= -0.009782139 Rot=63,35
||g|| = 0.020715980 Max(G)= -0.013196297 Rot=63,35
||g|| = 0.020835649 Max(G)= 0.011545400 Rot=65,35
||g|| = 0.025479718 Max(G)= -0.016634195 Rot=65,35

The program struggles with rotation 65-35. At this point we could play with the MaxRot
settings or switch convergence strategy. This will certainly require some trial and error.
In this particular case unrestricting the stepsize (MaxRot 5) does the job. In our
experience, the “orbstep SuperCI / switchstep DIIS” combination with a large level
shift is pretty robust. The scheme should be tried first when facing convergence
problems. Indeed convergence is achieved in 12 iterations. The remainder of this
calculation proceeds smooth. We report the final results with the ANO-RCC basis sets in
Table 4. The excitations energies have converged, but the results are still not optimal for
the doublet transition and the highest quartet state. The results should improve with
dynamical correlation.

Table 4.d-d transition energies for the [CrCl6]3- model complex in cm-1.

Excitations Exp.a SV ANO-RCC ANO-RCC ANO-RCC


ICE(13,15) DZP TZP FULL
99 MOs ICE(13,15) ICE(13,15) ICE(13,15)
141 MOs 263 MOs 780 MOs
4T 11900 12477 12466 12093 11992
2
2E 13545 19059 18609 18799 18770
2T 14180 19659 19275 19485 19467
1
4T
1(1) 18500 18442 19052 18734 18635
2T 20250 25452 25216 25391 25354
2
2A - 28354 27992 27810 27696
1
4T
1(2) 28500 28520 30942 30504 30369

25 G.L. Stoychev, A.A. Auer, and F. Neese, J. Chem. Theory Comput. 13, 554 (2017).
24

a O.S. Wenger, H.U. Güdel, J. Chem. Phys. 114, 5832-5841, 2001

6 [FeIV(O)(TMC)(MeCN)]2+ - covalent metal-ligand interactions and


the computation of MCD / Mössbauer
So far we have computed two Cr(III) model complexes, where the selection of the active
space was straight forward and the ligand orbitals could be identified easily. However,
this is not always the case. In the following example, we discuss strategies for systems
where both the metal and ligand orbitals are more delocalized. Further we illustrate
how to obtain the MCD spectra and Mössbauer parameters.

We chose a classic oxo-iron(IV) complex as an example, [Fe IV(O)(TMC)(MeCN)]2+ (TMC =


1,4,8,11-tetramethyl-1,4,8,11-tetraaza-cyclotetradecane), which has been
spectroscopically characterized to feature a triplet (S=1) ground state with a low-lying
quintet (S=2) state. Our major concern here is to compute reliable d-d transition
energies. Specifically, we demonstrate that the CAS(4,5) calculation, i.e. including only
the five metal d-orbitals and associated electrons, fails to predict accurate d-d transition
energies. To improve the description, one has to enlarge the active space by
incorporating ligand-based orbitals that strongly interact with the Fe d-orbitals. First, let
us perform a CASSCF calculation involving the five metal d-orbitals only (Figure 9). The
calculations are fairly large and should be executed on a cluster. For the same reason, we
use the RI approximation to speed-up the calculations.

Figure 9. CASSCF(4,5) natural orbitals of the [FeIV(O)(TMC)(MeCN)]2+ complex.

We start with a structure that was optimized at the B3LYP level and is further denoted
as “3_FeIV_FeO_TMC_B3LYP.xyz”. The origin is placed on the iron center. The
molecule is aligned so that the O2- ligand points in z direction, x- and y-axis point to
nitrogen ligands of the macrocyclic ring.
# xyz coordinated corresponding to 3_FeIV_FeO_TMC_B3LYP.xyz
* xyz 2 3
Fe 0.002998 -0.007344 0.001515
O -0.001116 -0.000277 1.630175
N 2.095349 -0.004573 0.097607
N 0.200201 2.113135 -0.058476
N -2.125671 0.063743 -0.055342
N -0.262696 -2.083062 0.099258
N 0.026185 -0.040306 -2.037309
C 2.379856 1.274701 0.818064
25

H 3.461553 1.463693 0.822094


H 2.049737 1.148041 1.850453
C 1.658766 2.417051 0.153944
H 2.100939 2.625540 -0.822455
H 1.760378 3.331135 0.750396
C -0.596737 2.654593 1.093961
H -0.443295 3.742414 1.111828
H -0.170268 2.236737 2.007879
C -2.096782 2.377084 1.031942
H -2.530900 2.871230 1.910687
H -2.547214 2.886931 0.172889
C -2.560266 0.924003 1.096929
H -2.195902 0.450687 2.010641
H -3.658578 0.907687 1.118175
C -2.609979 -1.345025 0.159227
H -3.529677 -1.331319 0.755585
H -2.872130 -1.758179 -0.816866
C -1.565990 -2.202307 0.823511
H -1.889548 -3.251614 0.831694
H -1.395726 -1.887484 1.854313
C 0.766571 -2.782101 0.930227
H 0.553252 -3.857800 0.873763
H 0.614242 -2.465696 1.965300
C 2.205440 -2.505017 0.533023
H 2.388950 -2.712817 -0.527779
H 2.833076 -3.216745 1.083469
C 2.659448 -1.112064 0.930432
H 2.362521 -0.920766 1.964913
H 3.753447 -1.034356 0.876226
C 2.786221 -0.008352 -1.220068
H 3.867100 0.090132 -1.060877
H 2.444812 0.817981 -1.841283
H 2.587643 -0.943763 -1.744176
C -0.183566 2.845719 -1.297287
H 0.021608 3.914507 -1.157466
H -1.238959 2.729346 -1.523970
H 0.403174 2.484394 -2.141228
C -2.804306 0.535065 -1.294551
H -3.890506 0.472572 -1.152725
H -2.524644 -0.097810 -2.136085
H -2.550379 1.564826 -1.526693
C -0.349394 -2.770800 -1.217076
H -0.583090 -3.830433 -1.055842
H 0.602407 -2.693001 -1.743597
H -1.127677 -2.329378 -1.837385
C 0.023227 -0.048464 -3.186690
C 0.016192 -0.058200 -4.631548
H -1.003427 0.121007 -4.993410
H 0.364485 -1.032499 -4.993982
H 0.679608 0.729029 -5.008050
*

Since we opt for the CAS(4,5) in the first run, we immediately start with PAtom as guess.
In order to calculate the d-d transition energies, the CAS(4,5) calculation is averaged
over ten triplet states using Def2-TZVP basis set. Once the calculation converges, we
perform a NEVPT2 correction on top the CASSCF wave function to see how dynamic
electron correlation affects the excitation energies. In the following input, we take a
shortcut and immediately start with the triple zeta basis and the PAtom guess. PAtom
automatically invokes a basis set projection from a singlet-zeta basis set. Hence, the
initial gradient will be large. In addition, we expect more covalent bonds in this example
(Figure 9), while PAtom produces very metal dominant orbitals. To avoid convergence
problems, we switch to a less aggressive scheme that is “orbstep SuperCI” and
“switchstep DIIS”. Note that the combination is particularly well suited to protect the
active space using level shifts.26
# d-d excitations with the CAS(4,5)

26 unless the occupation number is exactly 2.0 or 0.0


26

# def2-TZVP/C auxiliary basis (smaller compared to def2/JK)


!Def2-TZVP def2-TZVP/C TightSCF RI-NEVPT2 PAtom PAL8
%maxcore 4000
%casscf nel 4 # d-electrons
norb 5 # 3d-orbitals
mult 3 # triplet states
nroots 10 # calculate ten triplet states
trafostep ri # speed up integral trafo
orbstep SuperCI
switchstep DIIS
etol 1e-7 # reset etol: tightscf produces more accurate
# integrals and unnecessarily increases etol.
end

* xyzfile 2 3 3_FeIV_FeO_TMC_B3LYP.xyz

The CASSCF(4,5) calculation converges with the following composition of the wave
function for each state and the transition energies relative to the ground-state.
---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 3 NROOTS=10
---------------------------------------------

ROOT 0: E= -2235.4566946603 Eh
0.95940 [ 0]: 21100
0.01268 [ 3]: 20110
0.01100 [ 27]: 01120
0.00832 [ 7]: 12010
0.00658 [ 15]: 10210
ROOT 1: E= -2235.4145918856 Eh 1.146 eV 9240.5 cm**-1
0.72190 [ 6]: 12100
0.15436 [ 1]: 21010
0.06177 [ 17]: 10120
0.03150 [ 10]: 11110
0.02675 [ 25]: 01210
0.00250 [ 13]: 11011
ROOT 2: E= -2235.4122540846 Eh 1.209 eV 9753.6 cm**-1
0.72571 [ 9]: 11200
0.15270 [ 3]: 20110
0.06835 [ 12]: 11020
0.03276 [ 22]: 02110
0.01150 [ 15]: 10210
0.00278 [ 0]: 21100
ROOT 3: E= -2235.4035730634 Eh 1.446 eV 11658.8 cm**-1
0.96804 [ 10]: 11110
0.02283 [ 1]: 21010
0.00382 [ 25]: 01210
ROOT 4: E= -2235.3816933395 Eh 2.041 eV 16460.9 cm**-1
0.35487 [ 10]: 11110
0.35479 [ 1]: 21010
0.16455 [ 6]: 12100
0.08673 [ 25]: 01210
0.01405 [ 4]: 20101
0.01298 [ 17]: 10120
0.00406 [ 13]: 11011
0.00298 [ 16]: 10201
0.00293 [ 8]: 12001
ROOT 5: E= -2235.3809740494 Eh 2.060 eV 16618.8 cm**-1
0.27092 [ 7]: 12010
0.25558 [ 3]: 20110
0.23175 [ 15]: 10210
0.12533 [ 9]: 11200
0.08105 [ 22]: 02110
0.01255 [ 2]: 21001
0.00855 [ 12]: 11020
0.00514 [ 11]: 11101
ROOT 6: E= -2235.3776909500 Eh 2.150 eV 17339.3 cm**-1
0.66060 [ 10]: 11110
0.22014 [ 1]: 21010
0.06492 [ 25]: 01210
0.03331 [ 6]: 12100
0.00857 [ 4]: 20101
0.00555 [ 16]: 10201
0.00378 [ 8]: 12001
27

Let us have a close look at the definition of the wave function predicted by CAS(4,5)
calculation. The ground state wave function (ROOT 0) is dominated by the configuration
21100 (96%), corresponding to an electron configuration of (dxy)2(dxz)1(dyz)1(dx2-
y2) (dz2) . The first two excited states; ROOT 1 and ROOT 2 appear close to each other
0 0

with excitation energies 9240.5 and 9753.6 cm–1, respectively. These two excited states
are dominated by electronic configurations 12100 and 11200, representing the
transitions of dxy ® dxz and dxy ® dyz, respectively. Roots 3 and 6 are the dxy ® dx2-y2
transition. This is a shell-opening excitation, in which the number of the unpaired
electron increases from two to four. As a consequence, this single excitation gives rise to
five excited states in total. For more detailed discussion, we refer to the article of Ye et
al.27 The following higher energy excited states are mainly the excitations from dxz/yz to
the dx2-y2 (roots 4 and 5) and dz2 orbitals. The remaining transitions are mainly two-
electron excitations.

The successive NEVPT2 calculation on top of the CAS(4,5) wave function gives the
following excitation energies. In comparison with the experiment,28 the computed
excitation energies are significantly overestimated (Table 5).
-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0, MULT 3) = -2240.056906237 Eh -60955.047 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 1 3 0.063387 1.725 13911.8
2: 2 3 0.065484 1.782 14372.1
3: 3 3 0.093119 2.534 20437.2
4: 4 3 0.095106 2.588 20873.4
5: 5 3 0.099514 2.708 21840.8
6: 7 3 0.101362 2.758 22246.3
7: 6 3 0.103407 2.814 22695.2
8: 8 3 0.112200 3.053 24625.0
9: 9 3 0.159447 4.339 34994.7

Now we run a CASSCF calculation incorporating the ligand-based orbitals. In order to


design a balanced active-space that can deliver correct reference wave function, we need
to understand the bonding of the complex. The TMC ligand is a tetradentate ligand,
which form metal-ligand σ bond through the four equatorial N donors (σ-Neq). The oxo-
ligand forms very strong covalent bonds with the iron center involving one σ (σ-O) and
two π (2 x π-O) bonds. The axial MeCN also forms a σ bond (σ-Nax) with iron. Both σ-O
and σ-Nax simultaneously interact with Fe-3dz2 with a single bonding orbital that needs
to be included in the active space. Hence, the active space should contain four bonding
and the corresponding anti-bonding orbitals, in addition to the non-bonding Fe-3dxy
orbital. The complete active space is therefore constructed by twelve-electron
distributing over nine orbitals, CAS(12,9).

27Ye, S.; Xue, G.; Krivokapic, I.; Petrenko, T.; Bill, E.; Que, L., Jr; Neese, F., Chem. Sci. 2015, 6, 2909–2921
28Decker, A.; Rohde, J.-U.; Klinker, E. J.; Wong, S. D.; Que, L.; Solomon, E. I. J. Am. Chem. Soc. 2007, 129,
15983–15996.
28

Figure 10. CASSCF(12,9) natural orbitals of the [FeIV(O)(TMC)(MeCN)]2+ complex.

For the model complexes studied earlier, we could easily identify the ligand orbitals,
extend the active space and re-converge the calculation. Inspecting the doubly occupied
orbitals (range 0-96) in the converged CAS(4,5) calculation, we are not able to properly
select all four ligand orbitals depicted in Figure 10. The -ligand orbitals are entirely
missing. The majority of ligand orbitals do not have significant weight on the metal-d
orbitals e.g. Figure 11 illustrates how delocalized the canonical dx 2-y2-ligand orbital is.
Such orbitals would make a poor guess for the CASSCF(12,9) calculation and most
probably lead to convergence problems.

Figure 11. CAS(4,5) dx2-y2 ligand orbitals

---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 3 NROOTS=10
---------------------------------------------
29

From here, it is difficult to improve the CAS(4,5) orbitals. In our experience, for covalent
systems, QROs from DFT work well. Thus we proceed with QROs using the BP functional.
# def2/J RI auxiliary basis for pure DFT functionals
!BP def2-TZVP def2/J UNO pal8
%maxcore 4000
*xyzfile 2 3 3_FeIV_FeO_TMC_B3LYP_rotate-1.xyz

Figure 12 shows the ligand guess orbitals that we have selected. The QROs are not
“pure”, but at least the -ligand orbitals are present from the start.

Figure 12. Ligand orbitals selected QROs generated with the BP functional.

The CAS(12,9) output shows the familiar composition for the ground state and excited
state wave functions.
---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 3 NROOTS=10
---------------------------------------------

ROOT 0: E= -2235.6380115533 Eh
0.81233 [ 1469]: 222221100
0.04129 [ 1297]: 221122200
0.02851 [ 1019]: 211222101
0.02800 [ 1089]: 212121201
0.01427 [ 874]: 202221102
0.00717 [ 1462]: 222212010
0.00598 [ 1454]: 222210210
0.00581 [ 1442]: 222201120
0.00432 [ 737]: 122221110
0.00407 [ 684]: 122121210
0.00406 [ 614]: 121222110
0.00393 [ 774]: 201122202
0.00294 [ 232]: 022221120
0.00272 [ 466]: 112221111

A closer look to the configurations generated by CASSCF(12,9) calculation reveals that


the ground state (ROOT 0) is dominated by the configuration 222221100 (81%), which
corresponds to (σx2-y2)2(σz2)2(πxz/yz)4(dxy)2(π*xz/yz)2(σ*x2-y2)0(σ*z2)0. Following the same
30

route, one can assign the other excited states. The successive NEVPT2 calculation on top
of a CASSCF(12,9) wave function gives
-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0, MULT 3) = -2240.054966242 Eh -60954.995 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 1 3 0.052356 1.425 11490.8
2: 2 3 0.053857 1.466 11820.2
3: 3 3 0.058998 1.605 12948.6
4: 4 3 0.059058 1.607 12961.8
5: 8 3 0.059486 1.619 13055.6
6: 9 3 0.061982 1.687 13603.5
7: 5 3 0.062996 1.714 13826.1
8: 6 3 0.073581 2.002 16149.1
9: 7 3 0.075884 2.065 16654.5

Comparison of the major d-d transition energies at the CASSCF(4,5)/NEVPT2 and


CASSCF(12,9)/NEVPT2 levels to the experimental values below.
Table 5. NEVPT2 Excitation energies for the [FeIV(O)(TMC)(MeCN)]2+ complex

d-d transitions NEVPT2 NEVPT2 Exp.


CAS(4,5) CAS(12,9)
xz/yz → x2-y2 20873, 21840 11490, 11820 ~10600
xy → xz/yz 13911, 14372 13055, 13603 ~12900

Clearly the CASSCF(4,5)/NEVPT2 calculation significantly overestimates the transition


energies. While the transition energies predicted by CASSCF(12,9)/NEVPT2 agree
reasonably well with the experiment. This shows the importance of a balanced active
space and incorporation of the ligand orbitals. The observation simply reflects the fact
that a proper description of a covalent bond requires both bonding and anti-
bonding orbitals in the active space. This is a well-known protocol for studying bond-
breaking and bond-formation. Furthermore, in this situation, balancing the active space
improves the CASSCF convergence.

A remark on the character of the CASSCF orbitals:


For the current complex with covalent Fe–O bonds, the CASSCF(4,5) calculation predicts
ionic anti-bonding π-orbitals (88% Fe), whereas the CASSCF(12,9) calculation accurately
predicts covalent anti-bonding orbitals (62% Fe + 35% O). Judging the covalency based
on the unbalanced CASSCF(4,5) orbitals is dangerous!

6.1 Calculation of MCD spectra and Quadrupole splitting

Below is the input for a CASSCF/NEVPT2 calculation of the MCD spectrum. The
underlying physics are beyond the scope of this tutorial. The methodology is described
elsewhere. 29 Here, we briefly show how the MCD spectra can be computed with CASSCF.
A careful analysis of the MCD results can be found in the article of Ye et al.30

The MCD intensity is dominated by the spin-allowed transition. Hence, the previously
computed triplet d-d manifold should be sufficient. The computation involves

29 Ganyushin, D.; Neese, F., J. Chem. Phys. 2008, 128, 114117


30 Ye, S.; Xue, G.; Krivokapic, I.; Petrenko, T.; Bill, E.; Que, L., Jr; Neese, F., Chem. Sci. 2015, 6, 2909–2921
31

transitions from all microstates (all MS values). The corresponding keyword is


“NInitStates”. For the 10 non-relativistic triplet roots, there are 30 microstates (or
‘magnetic sublevels’ – these are the three MS sublevels for each of the ten triplet states).

!Def2-TZVP Def2-TZVP/C TightSCF RI-NEVPT2


!MOREAD
%moinp "3_FeIV_FeO_TMC_CAS_12-9_10T_c.gbw"
%pal
nprocs 8
end
%maxcore 4000

%casscf nel 12
norb 9
mult 3
nroots 10
TrafoStep ri
rel # enter into relativistic calculation
dosoc true # perform spin-orbit coupling
mcd true # perform the MCD calculation
NInitStates 30 # number of SOC state to account
# starts from the lowest state
NPointsTheta 10 # number of integration points for
NPointsPhi 10 # Euler angles
NPointsPsi 10
B 70000, 70000, 70000, 70000, 70000, 70000 # experimental
# magnetic field
# strength in Gauss
Temperature 2, 10, 20, 40, 60, 80 # experimental
# temperature (in K)
end
end

*xyzfile 2 3 3_FeIV_FeO_TMC_B3LYP.xyz

In the input file, the parameters magnetic field (B) and temperature are assigned in
pairs, i.e. B = 70000, 70000, 70000, . . . Temperature = 2, 10, 20. . . . The program
calculates MCD and absorption spectra for every pair. ORCA calculates the strength of
left circular polarized (LCP) and right circular polarized (RCP) transitions and prints the
transition energies, the difference between LCP and RCP transitions (intensity denoted
as C), and sum of LCP and RCP transitions (absorption intensity denoted as D), and C by
D ratio for every pairs of B and temperatures.

-----------------------------------------------------------
MCD Transitions
B = 70000.00 Gauss T = 2.000 K
-----------------------------------------------------------
C D C/D
-----------------------------------------------------------
0 -> 1 -0.00000 0.00154 -0.00004
0 -> 2 0.00000 0.00114 0.00005
0 -> 3 -0.00000 0.00000 -0.01153
0 -> 4 0.00000 0.00001 0.00050
0 -> 5 -0.00004 0.51146 -0.00008
0 -> 6 0.00004 0.33632 0.00012
0 -> 7 0.00000 0.03891 0.00000
0 -> 8 -0.00000 1.80573 -0.00000
0 -> 9 0.00000 0.39352 0.00001
0 -> 10 0.00000 32.45677 0.00000
0 -> 11 -0.00000 26.68497 -0.00000
0 -> 12 0.00001 0.87862 0.00001
0 -> 13 -0.00001 4.05381 -0.00000
0 -> 14 0.00000 0.01267 0.00001
0 -> 15 -0.00000 0.12560 -0.00000
0 -> 16 0.00000 1.51414 0.00000
0 -> 17 -0.00000 1.05308 -0.00000
0 -> 18 0.00001 0.94487 0.00001
0 -> 19 -0.00001 1.22514 -0.00000
0 -> 20 -0.00001 0.09362 -0.00008
32

0 -> 21 0.00001 0.09371 0.00008


0 -> 22 0.00000 0.05183 0.00000
0 -> 23 0.00000 1.06355 0.00000
0 -> 24 -0.00000 0.00739 -0.00001
0 -> 25 0.00000 0.00178 0.00004
0 -> 26 0.00000 0.06958 0.00000
0 -> 27 -0.00000 0.66457 -0.00000
0 -> 28 0.00000 0.16718 0.00000
0 -> 29 0.00000 0.91598 0.00000
0 -> 30 -0.00000 0.03829 -0.00001
0 -> 31 0.00000 0.72897 0.00000
0 -> 32 0.00000 0.33647 0.00000
0 -> 33 -0.00000 0.48522 -0.00000
0 -> 34 0.00000 0.42774 0.00000
1 -> 2 0.00000 0.00000 0.00005
1 -> 3 -0.00000 0.00002 -0.02590
1 -> 4 0.00000 0.00002 0.01821
1 -> 5 0.00000 0.00012 0.00000

In addition to the output, a successful calculation generates a series of files named like
input.1.casscf.mcd. The numbering identifies the magnetic field/ temperature
pair specified in the input. Since, we specified six pairs in the input, there should be files
numbered from 1 to 6. In case of NEVPT2, the files are named input.1.nevpt2.mcd
respectively. One can use orca-mapspc program to plot the predicted MCD spectra.
For details about orca_mapspc, please consult the orca manual. The keywords x0 and
x1 define the energy of the plot.

orca_mapspc input.1.nevpt2.mcd MCD -x04000 -x120000 -w2000

Here the interval for the spectra generation is set from 4000 cm–1 to 20000 cm–1, and the
line shape parameter is set to 2000 cm–1. If everything worked out fine, the program
prints a summary and produces a “.dat” file with the same name prefix.

Mode is MCD
Cannot read the *.mcd.inp file ...
taking the line width parameter from the command line
Number of peaks ... 66917
Start wavenumber [cm-1] ... 4000.0
Stop wavenumber [cm-1] ... 20000.0
Peak FWHM [cm-1] ... 2000.0
Number of points ... 1024

The dat file has 7 columns entries, where the first column is the energy. The next three
columns are the Gaussian convolution data points for the C , D and the ratio C/D. The last
three columns the discrete peaks (C,D and C/D). For a temperature of 2K, the resulting
MCD spectrum is depicted in Figure 13. The complete temperature depending MCD
spectra can be found in the article Ye et al.31

31 Ye, S.; Xue, G.; Krivokapic, I.; Petrenko, T.; Bill, E.; Que, L., Jr; Neese, F., Chem. Sci. 2015, 6, 2909–2921.
33

Figure 13. MCD spectra for a temperature of 2K.

6.2 Mössbauer Parameters


The example input below shows how to calculate the Mössbauer parameters for the iron
complex using the CASSCF wave function. The orca_eprnmr module typically applies
to single reference methods. The Mössbauer parameters are an exception in this
program. Here we directly use the previously converged CAS(12,9) wave function and
add the necessary keywords to the %eprnmr block. Note that the %eprnmr block must
be placed below the coordinate block. As a reminder, the quadrupole splitting is a
ground state property. It is thus important to restrict the calculation to nroots=1 to
get the ground-state density. Do not use the state-averaged density for the
computation of Mössbauer Parameters! NEVPT2 is omitted, since the NEVPT2 density is
not yet available (it is important to understand that the NEVPT2 procedure only corrects
the energy, not the wave function and hence also not the density!).
!Def2-TZVP Def2-TZVP/C TightSCF
!MOREAD
%moinp "CAS_9_10Roots.gbw" # converged CAS(12,9) with 10 triplet roots.
%pal
nprocs 8
end
%maxcore 4000

%casscf nel 12
norb 9
mult 3
nroots 1
TrafoStep ri
end

*xyzfile 2 3 3_FeIV_FeO_TMC_B3LYP.xyz

%eprnmr
Nuclei = all Fe {fgrad, rho}
end
34

The output file should contain the following lines at its end, where you obtain the
calculated quadrupole splitting (Delta-EQ) directly and the RHO(0)value (the electron
density at the iron nucleus).

-----------------------------------------
ELECTRIC AND MAGNETIC HYPERFINE STRUCTURE
-----------------------------------------

-----------------------------------------------------------
Nucleus 0Fe: A:ISTP= 57 I= 0.5 P= 17.2798 MHz/au**3
Q:ISTP= 57 I= 0.5 Q= 0.1600 barn
-----------------------------------------------------------
Tensor is right-handed.

Raw EFG matrix (all values in a.u.**-3):


-0.1898 0.0177 -0.1423
0.0177 -0.1922 0.1536
-0.1423 0.1536 0.3820

V(El) -0.2058 -0.1844 0.3902


V(Nuc) 0.0325 -0.0909 0.0585
---------- ---------- ----------
V(Tot) -0.1733 -0.2754 0.4487
Orientation:
X 0.7327196 0.6485519 0.2061613
Y 0.6805305 -0.6980558 -0.2227023
Z -0.0005219 0.3034774 -0.9528385

Moessbauer quadrupole splitting parameter (proper coordinate system)


e**2qQ = 16.890 MHz = 1.456 mm/s
eta = 0.228
Delta-EQ=(1/2{e**2qQ}*sqrt(1+1/3*eta**2) = 8.517 MHz = 0.734 mm/s
RHO(0)= 11586.879407946 a.u.**-3

The calculated quadrupole splitting (0.734 mm/s) agrees very well with the experiment
value of 1.24mm/s,32 which again credence our chosen active space.

7 [Co(SH)4]2- - Optical and Magnetic properties


7.1 Electronic Structure
In ideal Td geometry, the tetracoordinate Co(II) complexes possess a 4A2 ground state,
with a half-filled t2 subshell. The important single excitations within the metal d-shell
are those from the doubly occupied e-orbitals (dz2 and dx2−y2) to the singly occupied t2-
set (dxy, dxz, and dyz). These excitations give rise to two quartet-excited states (4T1 and
4T ). Under conditions favoring further symmetry lowering, there will be further
2
splitting of the x, y, z components of the T states, such that T2x≡1-Ex, T2y≡1-Ey, T2z≡Bz
and T1x≡2-Ex, T1y≡2-Ey, T1z≡Az in ~ S4 symmetry.

32 Rohde, J.-U.; In, J.-H.; Lim, M. H.; Brennessel, W. W.; Bukowski, M. R.; Stubna, A.; Münck, E.; Nam, W.; Que,
L. Jr. Science 2003, 299, 1037–1039.
35

Figure 14. The metal d-based MOs of the model complex [Co(SH)4] 2-. Final state term symbols arising from
single excitations are analyzed under approximate S 4 symmetry. The indicated orbital occupation pattern
refers to the 4A2 ground state.

7.2 Setting up the CASSCF/NEVPT2 calculation


A general input for performing optical and magnetic properties calculations within the
CASSCF/NEVPT2 methodology for the pseudo tetrahedral Co II (S=3/2) model complex
(Figure 14) is provided below:
! def2-TZVP def2-TZVP/C PAtom PAL4

%casscf nel 7
norb 5 #7 electrons in 5 d orbitals
nroots 10,35
mult 4,2 # 10 quartet and 35 doublet states
trafostep RI
#------------------------------------------------
PTMethod SC_NEVPT2 #Perform the SC-NEVPT2 correction
#------------------------------------------------
rel #flag for relativistic properties
printlevel 3 #Control the amount of printing
dosoc true #Do the SOC calculation
#-----------------------------------------------
mcd true # Request the MCD calculation
NInitStates 28 # Number of Donor SOC states
# for the ABS and MCD spectra evaluation
NPointsTheta 10 # Number of integration point for
NPointsPhi 10 # Euler angles
NPointsPsi 10 #
B 5000 # Experimental Magnetic field (in Gauss)
Temperature 10 # Experimental temperature (in K)
#-----------------------------------------------
gtensor true # Request the G-tensor Calculation
#-----------------------------------------------
dtensor true # Request the ZFS-tensor Calculation
#(default if dosoc true)
#-----------------------------------------------
end
end

* xyz -2 4
Co 0.000089000 0.000270000 0.000180000
S -1.194914000 1.441126000 -1.471908000
36

S -1.394869000 -1.170677000 1.534517000


S 1.120414000 -1.419512000 -1.549797000
S 1.467145000 1.145600000 1.486755000
H 2.291750000 1.687013000 0.542215000
H -1.704914000 2.280708000 -0.523177000
H -2.260149000 -1.693145000 0.616185000
H 1.675644000 -2.271185000 -0.637921000
*

In this protocol the minimal active space (3d orbitals) is chosen to be a CAS(7,5). As seen
in the Section 5, CASSCF orbitals restricted to the 3d-metal orbitals are very ionic.
In this example the converged CASSCF orbitals have more than 89% metal character
according to the Loewdin population analysis. Thus, the PAtom guess is the ideal choice
for such a setup.

NEVPT2 is performed on top of the converged state averaged CASSSCF wave function. It
should recover a major part of the dynamic electron correlation between the ground and
the excited states. The calculation is carried out on the basis of all 10 roots for the
quartet states (arising from the 4F and 4P terms of Co2+) together with 35 doublet roots
(arising from the 2G, 2H, 2F, 2P, 2D terms of Co2+). Furthermore, in order to avoid
unrealistic description of the ground state due to the state averaging the energetically
higher second 2D term (accounting for the rest 5 doublet roots) is excluded.

The output contains the following CASSCF and NEVPT2 transition energies.
-----------------------------
SA-CASSCF TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0 ,MULT 4) = -2974.012986842 Eh -80927.008 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 1 4 0.003979 0.108 873.3
2: 2 4 0.015774 0.429 3462.0
3: 3 4 0.015782 0.429 3463.7
4: 4 4 0.019875 0.541 4362.0
5: 5 4 0.019882 0.541 4363.5
6: 6 4 0.024591 0.669 5397.1
7: 0 2 0.088630 2.412 19452.0
8: 1 2 0.088764 2.415 19481.5
9: 2 2 0.089205 2.427 19578.2
10: 3 2 0.089210 2.428 19579.3
11: 4 2 0.089837 2.445 19716.9
12: 5 2 0.090128 2.453 19780.8
13: 6 2 0.093332 2.540 20483.9
---------------------------------------------

-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0, MULT 4) = -2975.223602565 Eh -80959.950 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 1 4 0.005873 0.160 1288.9
2: 2 4 0.023207 0.631 5093.3
3: 3 4 0.023218 0.632 5095.8
4: 4 4 0.029850 0.812 6551.3
5: 5 4 0.029857 0.812 6552.9
6: 6 4 0.035846 0.975 7867.4
7: 2 2 0.083339 2.268 18290.9
8: 3 2 0.083349 2.268 18293.0
9: 0 2 0.083708 2.278 18371.9
---------------------------------------------
37

Inspection of the states composition shows that the first 6 quartet excited states
correspond to the z and xy components of the 4T2 and 4T1 final states, respectively
(Figure 14).
---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 4 NROOTS=10
---------------------------------------------

ROOT 0: E= -2974.0129868416 Eh
0.99920 [ 9]: 22111
ROOT 1: E= -2974.0090077771 Eh 0.108 eV 873.3 cm**-1
0.99920 [ 8]: 21211
ROOT 2: E= -2973.9972125969 Eh 0.429 eV 3462.0 cm**-1
0.54004 [ 4]: 12121
0.37209 [ 7]: 21121
0.05208 [ 1]: 11212
0.03120 [ 3]: 12112
0.00362 [ 6]: 21112
ROOT 3: E= -2973.9972050318 Eh 0.429 eV 3463.7 cm**-1
0.53816 [ 3]: 12112
0.37324 [ 6]: 21112
0.05251 [ 2]: 11221
0.03144 [ 4]: 12121
0.00373 [ 7]: 21121
ROOT 4: E= -2973.9931121667 Eh 0.541 eV 4362.0 cm**-1
0.66405 [ 2]: 11221
0.20722 [ 3]: 12112
0.05145 [ 1]: 11212
0.04247 [ 6]: 21112
0.02355 [ 7]: 21121
0.01126 [ 4]: 12121
ROOT 5: E= -2973.9931051642 Eh 0.541 eV 4363.5 cm**-1
0.66485 [ 1]: 11212
0.20578 [ 4]: 12121
0.05167 [ 2]: 11221
0.04300 [ 7]: 21121
0.02354 [ 6]: 21112
0.01115 [ 3]: 12112
ROOT 6: E= -2973.9883958910 Eh 0.669 eV 5397.1 cm**-1
0.70304 [ 0]: 11122
0.29695 [ 5]: 12211
ROOT 7: E= -2973.9179925415 Eh 2.585 eV 20848.8 cm**-1
0.70300 [ 5]: 12211
0.29693 [ 0]: 11122
ROOT 8: E= -2973.9171526473 Eh 2.608 eV 21033.2 cm**-1
0.50849 [ 7]: 21121
0.23066 [ 1]: 11212
0.21146 [ 4]: 12121
0.04933 [ 6]: 21112
ROOT 9: E= -2973.9171455718 Eh 2.608 eV 21034.7 cm**-1
0.50777 [ 6]: 21112
0.23074 [ 2]: 11221
0.21226 [ 3]: 12112
0.04913 [ 7]: 21121

7.3 Optical spectra

Based on these excited states several spectra including Absorption, CD (Circular


Dichroism) and SOC corrected Absorption spectra (provided that relativistic calculations
have been requested) are generated. These spectra can be extracted by using the
orca_mapspc utility. The orca_mapspc program applies Gaussian/Lorenzian type
line shape functions to the calculated transitions with a user-defined “full width at half
maximum” (FWHM) value. One has to provide some information for the program such as
the name of the output file, the type of the desired spectrum and the energy range.

# The lines below are shell commands that call orca_mapspc directly
# myfile.out is the output of the previous calculation.

orca_mapspc myfile.out ABS –x05000 –x125000 –w1500 –n10000


38

orca_mapspc myfile.out SOCABS –x05000 –x125000 –w1500 –n10000


orca_mapspc myfile.out CD –x05000 –x125000 –w1500 –n10000

The output contains CASSCF and NEVPT2 results. The orca_mapspc program
processes both results. In the following we focus on the NEVPT2 results. In Figure 15
(top) the respective Absorption and SOC corrected Absorption spectra for [Co(SH)4]2-
are presented in the energy range 0 -10000 cm-1. As can be seen the Absorption
spectrum is dominated by electron excitations into the 1 - 4 Ex , y , 2 - 4 Ex, y and 4 Az (Figure
15, top left and Figure 1). Moreover, switching on the SOC effects a more complicated
spectral pattern is observed owning to the nature of the contributing Ms states (Figure
15, top right). This can modify the intensity, the shape as well as the energy distribution
of the spectrum. In general, this is expected to be more pronounced in higher energy
areas, which are dominated, by states of different multiplicities (Figure 15, bottom left
and right).

Figure 15. NEVPT2 Visible Absorption and SOC corrected Absorption spectra of [Co(SH)4]2-

7.4 G-tensor
39

CASSCF provides two methodologies that can deliver information about the g-tensors.
These are:

· The Effective Hamiltonian method.33 which provides information about the


principal g values and it is hence valid for every spin case (Kramers or non-
Kramers systems)
· The quasi-degenerate perturbation theory (QDPT) is valid for an individual
Kramers doublet. In fact, for a system with odd electrons (non-integer spin cases
or Kramers systems) the doubly degenerate eigenvalues obtained from the QDPT
procedure represent Kramers pairs. Hence one can obtain information about the
effective g values of an isolated Kramers doublet assuming pseudo spin S=½.

Let us first inspect the g values estimated by the Effective Hamiltonian method.

----------------------------------------------
ELECTRONIC G-MATRIX FROM EFFECTIVE HAMILTONIAN
----------------------------------------------

g-matrix:
2.243559 -0.000189 0.018262
-0.000120 2.242967 -0.006258
0.018210 -0.006289 2.995835

g-factors:
2.242915 2.243117 2.996329 iso = 2.494120

g-shifts:
0.240596 0.240798 0.994009 iso = 0.491801

Eigenvectors:
X 0.012637 0.999627 0.024214
Y 0.999888 -0.012438 -0.008333
Z 0.008029 -0.024317 0.999672

# The g-factors are square roots of the eigenvalues of gT*g


# Orientations are the eigenvectors of gT*g

The output shows total g-matrix followed by the three principal components and their
orientation, where the orientation (eigenvectors). Note that the eigenvectors are printed
as column vectors – see color coding. The orientation can also be used to rotate the
molecule into the eigenframe of the g-tensor – simply multiple the xyz coordinates with
the orientation tensor. In this example, we obtain gz=2.99 and gx,y=2.24.

Furthermore knowing the ratio of the g-values (gz/gx,y=1.3) and the orientation
(eigenvectors) of the g tensor in the molecular axis frame one can be visualize the g-
tensor components (See Figure 16).

33 Atanasov, M.; Aravena, D.; Suturina, E.; Bill, E.; Maganas, D.; Neese, F.; Coord. Chem.. Rev. 2015, 289-290
(0), 177-214
40

Figure 16. Orientation of the g tensor components in the molecular frame

In general, the g-values are only influenced by spin-allowed (spin-conserving)


transitions. Analysis from the individual states reveals the only states that contribute are
those that can interact via spin-orbit coupling with the ground state. Hence only the first
three states contribute. According to the expectations from the LFT,34 the g values are
mainly dominated by single electron excitations into the T 2 states: 4A2→4Ex (dx2-y2→dxz),
4A →4Ey (d
x2-y2→dyz) and A2→ A2 (or B )(dx2-y2→dxy). The largest positive g-shift is
4 4 z 4 z
2
calculated for the gz parameter, due to the strong SOC effect of the excited 4Bz on the 4A2
ground state.

Additionally, the effective g'-tensors, arising from a particular Kramers doublet serves a
diagnostic tool for the composition of the ground state Kramers doublet and thus
possibly the sign of zero field splitting D.
--------------
KRAMERS PAIR 1
--------------

g-factors:
0.003963 0.005248 8.870597 iso = 2.959936

--------------
KRAMERS PAIR 2
--------------

g-factors:
3.374535 4.384133 4.385082 iso = 4.047917

As can be seen the ground state Kramers doublet represents an axial system with
=
g z' 8.9, g x' , y < 0.1 . On the other hand the second Kramers pair represents a rhombic
system with=
g z' 3.4 =
g x' , y 4.4 . This indicates that the ground state Kramers doublet is
composed of=
M s ± 3 / 2 . In fact by inspecting the composition of the calculated SOC
states this is exactly what is observed:

34 Neese, F.; Solomon, E. I., Magnetoscience - Molecules to Materials. Miller, J.S.;Drillon, M. ed.; 2003; Vol.
IV, p 345
41

Eigenvectors:
Weight Real Image : Block Root Spin Ms
STATE 0: 0.0000
0.819191 -0.099210 0.899638 : 0 0 3/2 3/2
0.169283 0.408961 0.045101 : 0 1 3/2 3/2

STATE 1: 0.0000
0.819191 0.876339 -0.226321 : 0 0 3/2 -3/2
0.169283 0.102883 0.398370 : 0 1 3/2 -3/2

STATE 2: 130.5095
0.873822 -0.927664 -0.115162 : 0 0 3/2 1/2
0.029847 -0.023661 0.171136 : 0 1 3/2 1/2
0.061898 -0.011361 -0.248534 : 0 0 3/2 -1/2

STATE 3: 130.5095
0.061898 -0.244423 -0.046429 : 0 0 3/2 1/2
0.873822 -0.017317 0.934624 : 0 0 3/2 -1/2
0.029847 -0.172762 -0.000803 : 0 1 3/2 -1/2

7.5 D-tensor

If the calculation of the D-tensor is requested, (e.g. D tensor true), both the 2nd Order PT
approach35 as well as the Effective Hamiltonian Method are used to evaluate the sign and
the value of the zero field splitting D (note that if the SOC flag is on and the total spin of
the system is S>1/2 this information is printed by default). In general both approaches
are adequate in calculating the D-tensor18 (and references there in). In the case of
[Co(SH)4]2- and in accordance with the discussion above both approaches predict an
axial system with E/D~0 and negative D. However the 2nd order PT approach
overestimates D by a factor of 1.5. As shown elsewhere18,36, in the presence of low lying
excited states (< 1000 cm-1), which contribute through SOC with the ground state, the
2nd order PT methodology is not valid for the calculation zero-field splitting (ZFS). Under
such circumstances, the D values must be analyzed carefully.
--------------------------------------------
ZERO-FIELD SPLITTING
(2ND ORDER SPIN-ORBIT COUPLING CONTRIBUTION)
--------------------------------------------


Direction X=1 Y=2 Z=0
D = -101.374638 cm-1
E/D = 0.000096

--------------------------------------------------------
ZERO-FIELD SPLITTING
(EFFECTIVE HAMILTONIAN SPIN-ORBIT COUPLING CONTRIBUTION)
--------------------------------------------------------

Direction X=1 Y=2 Z=0


D = -65.254746 cm-1
E/D = 0.000096

Furthermore, analysis of the individual contributions to the D-tensor shows that once
again the major contributions to the ZFS originates from 4T2 states dominated by
4A →4Ex (d
x2-y2→dxz), A2→ E (dx2-y2→dyz) and 4A2→4A2z(or 4Bz)(dx2-y2→dxy) single
4 4 y
2
electron excitations. Moreover, significant contributions are observed from the
respective 2T2 states.

35 Neese, F. and Solomon, E. I., Inorg. Chem. 1998, 37, 6568


36 Maganas, D.; Sottini, S.; Kyritsis, P.; Groenen, E. J. J.; Neese, F., Inorg. Chem. 2011, 50 (18), 8741-8754
42

Individual contributions to D-tensor:


Block Mult Root D E
0 4 0 0.000 0.000
0 4 1 -103.443 -0.000
0 4 2 15.503 -15.344
0 4 3 15.489 15.328

1 2 6 6.678 -0.000
1 2 7 -2.391 1.707
1 2 8 -2.390 -1.707

8 Cu-dimer J-Couplings with NEVPT2, DLPNO-NEVPT2 and MRCI


The Copper dimer [Cu(tmeen)(OH)]2Br2 shown in Figure 17 is a toy system, where a
number of theoretical methods have been exercised. In this section, we study the singlet-
triplet splitting. The two Cu2+ atoms of the dimer have a [Ar]3d9 electronic configuration
and therefore have one magnetic electrons each. This Cu dimer which we shall call
Cu2(OH)2 henceforth, has an anti-ferromagnetic coupling constant of Jexp=-509cm-1
between the two magnetic centers.37 In the following, we use the crystal structure with
optimized hydrogen positions (cu2.xyz). The minimal active space describing the
singlet-triplet splitting is a CAS(2,2). The quality of ab initio results strongly depends on
the methodology and it has been shown that such a minimal CAS does not work for
multireference perturbation theory (e.g. NEVPT2).38

Figure 17: The two (localized) magnetic orbitals of Cu2(OH)2.

For the purpose of the tutorial, we use only small basis set (SV) and omit relativistic
effects. Nevertheless, the example with ~500 basis function is rather big. It is a perfect
opportunity to discuss the usage of the RI approximation in the context of
multireference methods (Section 8.3). Note that Cu complexes are known to cause
convergence problems due to the fully packed 3d shells. Therefore, we also discuss some
common issues here.

8.1 CASSCF(2,2) and NEVPT2

As discussed in the previous sections, for active spaces consisting of the “ionic” metal
orbitals, PAtom is the guess of choice. To speed-up the CASSCF calculations, we use the
RI approximation for the integral transformation (“trafostep RI”) as well as the Fock
operator build (RIJK or RIJCOSX)39. The RIJCOSX is more efficient for larger molecules

37 Bill J Cole, J. Chem. Phys. 53, 4718–19 (1970)


38 C.J. Calzado, C. Angeli, D. Taratiel, R. Caballol, and J.-P. Malrieu, J. Chem. Phys. 131, 44327 (2009)
39 “!RIJK” and “!RIJCOSX” automatically set “Trafostep RI” in CASSCF.
43

(more than 1000 basis functions). For not too big molecules, we prefer the RIJK
approximation that is numerically more stable.40 Both approximations reduce the
computational time considerably. The resulting orbitals gradients are very similar!
Hence if the accuracy with RI is not sufficient, we can still use the resulting orbitals as a
guess for the actual non-RI calculation. This saves computational time.
! def2-SVP RIJK conv def2/JK PAtom PAL8
%maxcore 3000 # 3GB of allowed memory usage per process
%casscf nel 2
norb 2
mult 3,1
# By default equal weights are given for each multiplicity block
nroots 1,1
# Localize the orbitals for an easier understanding of the wave function
actorbs locorbs
end

* xyzfile 2 3 cu2.xyz

Below is the output of the CASSCF(2,2) state averaged calculation, where we compute
one root for each multiplicity. In the following we use equal weights for both
multiplicities. As discussed in reference 38, “good” orbitals are crucial for the correct
prediction of the singlet-triplet splitting. It is possible to improve the results by
including the next singlet state and shifting the weight towards the singlet (e.g. 70%). In
our opinion, this is a door that should not be opened as it questions the predictive power
of the method. Instead we prefer to improve the methodology systematically by
including dynamical correlation (vide infra).
---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 3 NROOTS= 1
---------------------------------------------

ROOT 0: E= -4118.2116939264 Eh
1.00000 [ 0]: 11

---------------------------------------------
CAS-SCF STATES FOR BLOCK 2 MULT= 1 NROOTS= 1
---------------------------------------------

ROOT 0: E= -4118.2119500614 Eh
0.99939 [ 1]: 11

-----------------------------
SA-CASSCF TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0 ,MULT 1) = -4118.211950061 Eh -112062.244 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 0 3 0.000256 0.007 56.2

In order to calculate the Heisenberg coupling constant J defined by the Hamiltonian


equation: = − ( ∙ ), one needs to know the energy difference between the singlet
and triplet states. The energy difference, printed in the last line above, directly
corresponds to the coupling constant, i.e. ∆ = . At the CASSCF level, the ground state
is predicted correctly as anti-ferromagnetically coupled. However, the value of J=-56
cm-1 is far away from the experimental findings.

40 RIJCOSX introduces an integration grid (GridX). Poor grid settings may lead to convergence problems.
44

Next, we compute the strongly contracted NEVPT2 correction using the converged
orbitals.
!def2-SVP RI-JK conv def2/JK moread RI-NEVPT2 PAL8
%moinp “cas2.gbw” # converged CASSCF(2,2)
%maxcore 3000 #more memory
# CAS 2 electrons in 2 orbitals
%casscf nel 2
norb 2
mult 3,1
# Keep the initial shape of the active orbitals (localized in this example)
# This is of no importance for the NEVPT2 and is shown here for fun.
actorbs unchanged

end
* xyzfile 2 3 cu2.xyz

Dynamical correlation indeed improves the energy separation. NEVPT2 only recovers
~23% of the experimental J value. The results should further improve with a bigger
basis set, but this will not be a game-changer.
-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0, MULT 1) = -4121.682751137 Eh -112156.690 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 0 3 0.000535 0.015 117.3

8.2 CASSCF(2,2) DDCI3 - the game changer

For the CAS(2,2) and the given basis, the potentially best results are obtained with the
difference dedicated MRCI (DDCI3). The approach is essentially MRCI with singles and
doubles, where the two-hole two-particle type excitations are omitted.41 The approach is
computationally attractive for excitation energies. The MRCI, as implemented in the
orca_mrci module, is an uncontracted MRCI, where CFSs constructed from singly
and doubly excited CFGs with respect to all reference CFGs. Consequently, the number
of CFS grows rapidly with the size of reference space. Moreover, the integrals must
be kept in memory, which limits the size of the applicable molecules. In this particular
example, the calculation is feasible, since there are only 3 reference configurations (20,
02 and 11).

We again start with the converged CASSCF. Note that we set the “allowrhf” flag with
“noiter”, otherwise the program will re-optimize the orbitals using the Hartree Fock
method (single reference!). The following input uses the minimal set of input
parameters. The orca_mrci by default truncates the CI space to keep the calculation
manageable. The most important parameters are TPre and TSel for the perturbation
selection.42 The parameters can impact the computed J values and should be validated
for a rigorous study.43

# nopop disables the MRCI population analysis. This can take substantial
# time when computing many roots.

41 J. Miralles, O. Castell, R. Caballol, and J.-P. Malrieu, Chemical Physics 172, 33 (1993).
42 F. Neese, T. Petrenko, D. Ganyushin, and G. Olbrich, Coordination Chemistry Reviews 251, 288 (2007).
43 The perturbation selection also depends on the representation of the active orbitals.
45

! def2-SVP MOREAD PAL4 noiter allowrhf nopop


%moinp "cas2.gbw" #converged CASSCF(2,2) orbitals
%maxcore 6000 # 6GB of memory usage per process
%mrci citype mrddci3

# preselection cut-off
TSel 1e-8 # tighter than default TSel 1e-6
#TPre 1e-4 #default

# singlet multiplicity block


newblock 1 *
refs cas(2,2) end #reference wave function
end
# triplet multiplicity block
newblock 3 *
refs cas(2,2) end #reference wave function
end
davidsonopt none # no davidson correction

# create natural orbitals (not necessary)


natorbiters 1 # a value of 1 produces average nat. orbitals
# and does a second MRCI calculation with the
# new orbitals.
end

* xyzfile 2 3 cu2.xyz

In the input snippet above, we requested natural orbitals. The program creates a file
“.mrci.nat”, which is a regular gbw-file that can be read by ORCA. As shown in the review
by Calzado et al,38 the CAS(2,2) NEVPT2 results dramatically improve with natural
orbitals from DDCI3. If you are not interested in natural orbitals, drop the respective
keyword line.

In this calculation we have switched off the population analysis (nopop). This avoids the
construction of MRCI densities, which saves time and computer resources.

Note that we have disabled the Davidson type size-consistency corrections. Size-
consistency errors have a strong influence on the computed exchange coupling
parameters. The Davidson correction strongly relies on the weight of the reference wave
function and is unreliable for small reference weights. DDCI3 is subject to size-
consistency errors, but much less than the canonical MRCI (singles doubles). Hence, the
results without size-consistency corrections should be reasonable.

The output consists of two CI blocks, one for the triplet state and one for the singlet
state. Since local orbitals have been used, the analysis can be done in terms of metal
centered local orbitals. Looking first at the J value or the energy difference44 between the
singlet and triplet state we see a huge improvement of -431 cm-1:
TRANSITION ENERGIES
-------------------

The lowest energy is -4118.426067758 Eh

State Mult Irrep Root Block mEh eV 1/cm


0 1 -1 0 0 0.000 0.000 0.0
1 3 -1 0 1 1.963 0.053 430.8

44 The transition energy in part stems from the actual DDCI3 calculation and in part of the perturbative
correction of the unselect CSF. This energy correction can be substantial and is printed as “E(unsel)” for
each state.
46

The MRCI program prints the weight of the different CFG's with large contributions on
top of the CAS(2,2) space. The wave function after the DDCI3 calculation is shown below
first for the triplet and then for the singlet state:

STATE 0: Energy= -4118.424104868 Eh RefWeight= 0.9003


0.9003 : h---h---[11]
0.0045 : h---h 84[21]
0.0043 : h---h100[21]
0.0049 : h---h101[21]
0.0045 : h---h 84[12]
0.0041 : h---h100[12]
0.0050 : h---h101[12]

The composition of the singlet wave function is shown below:


STATE 0: Energy= -4118.426067758 Eh RefWeight= 0.8945
0.0033 : h---h---[20]
0.8880 : h---h---[11]
0.0033 : h---h---[02]
0.0030 : h---h 82[21]
0.0058 : h---h 84[21]
0.0056 : h---h100[21]
0.0037 : h---h101[21]
0.0030 : h---h 82[12]
0.0058 : h---h 84[12]
0.0055 : h---h100[12]
0.0038 : h---h101[12]

The two outputs above can be interpreted as following:


· The configurations (CFGs) are given for the CAS space selected i.e. (2,2).
Therefore, all the CFGs contain two orbitals [xy] where x and y are the
occupations for the two orbitals.
· For the triplet state, the wave function is still strongly dominated by the [11]
CFG which accounts for almost 90 percent of the wave function as expected.
· For the singlet state, on the other hand, the weight of the [11] CFG goes from
99.9% (CASSCF) to 88.8%. This significant decrease in the [11] CFG is
contrasted by an increase from 0.03% to 0.33% of the [20] and [02] i.e. "Ionic"
CFGs. This increase of the "Ionic" CFGs by more than a factor of 10 explains the
increase in the J coupling constant.
· Dominant CFG's outside the reference CAS(2,2) space are the 1h CFGs. They
correspond to an excitation of 1 electron from the inactive orbitals to the active
orbitals. The h84[12] CFG represents such a CFG.

8.3 RI approximation for CASSCF, NEVPT2, DLPNO-NEVPT2 and MRCI


The RI-Approximation is a good choice for relative energies. Nevertheless, one
might wonder how big the influence on sensitive properties such as the exchange
coupling is. Table 6 documents the NEVPT2 J values computed with the following
approximations:

· “Trafostep RI” in conjunction with a /JK auxiliary basis set


(corresponds to AuxC=def2/JK)
!def2-svp def2/JK ...
%casscf TrafoStep RI
47

· “Trafostep RI” and !RIJK in conjunction with a /JK auxiliary basis set corresponds
to AuxC=def2/JK, AuxJK=def2/JK)
!def2-svp RIJK def2/JK ...

· “Trafostep RI” and !RIJCOSX in conjunction with a /JK auxiliary basis set
corresponds to AuxC=def2/JK, AuxJ=def2/JK)
!def2-svp RIJCOSX def2/JK ...

· “TrafoStep RI” and !RIJCOSX providing a /J and /C auxiliary basis sets


(corresponds to AuxC=def2-TZVP/C, AuxJ=def2/J). This is much more efficient!
!def2-svp RIJCOSX def2-TZVP/C def2/J

As a reminder, starting with ORCA 4.1, the program internally uses auxiliary basis slots
(AuxC, AuxJK and AuxJ). In general, the AuxJK slot is used in conjunction with the RIJK
approximation, whereas the AuxJ slot is associated with the RIJCOSX approximation. All
other RI approximated integrals (CASSCF orbital gradient, CASSCF orbital Hessian,
NEVPT2, MRCI) use the auxiliary basis defined in the AuxC slot. The auxiliary basis sets
are automatically assigned to their respective slots, unless the %basis is explicitly
invoked to set auxiliary basis sets.

Table 6. Exchange coupling [cm-1] computed with various levels of RI approximation. aRIJCOSX with a single
auxiliary basis (AuxC=AuxJ=def2/JK). bRIJCOSX with AuxC=def2-TZVP/C and AuxJ=def2/J assigned separately.

NORI (exact) RI RIJK RIJCOSXa RIJCOSXb


CASSCF -56.2 -56.2 -56.2 -56.3 55.4
NEVPT2 -117.0 -117.3 -117.3 -116.9 115.7
FIC-NEVPT2 -124.9 -124.9 -124.9 -125.2 123.4
DLPNO-NEVPT2 -125.1 -125.0 -127.2 123.8
DDCI3 -430.8 -430.7 -435.9 -429.5 -

All of the RI approaches speed-up the calculation without a noticeable loss of accuracy.
RIJK is generally recommended as long as you can store the three-index integrals.
Orbital gradients computed with RIJCOSX are slightly different (grid/gridx dependent).
Hence for CASSCF optimization we prefer the RIJK approximation or the simpler
“trafostep RI”, which are almost spot-on with the “exact” results.

For historical reasons, we used the term NEVPT2, although the correct terminology is
strongly contracted NEVPT2 (SC-NEVPT2). The SC-NEVPT2 is computationally
attractive and has been in a workhorse in our own group. The difference between SC-
NEVPT2 and the fully internally contracted NEVPT2 (FIC-NEVPT2 aka PC-NEVPT2) is
small.45,46

For larger molecules (more than 80 atoms), the fastest NEVPT2 approach is the
DLPNO-NEVPT2 recently developed by Guo et al.47 It is the local-correlation extension

45 I. Schapiro, K. Sivalingam, and F. Neese, J. Chem. Theory Comput. 9, 3567 (2013).


46 R.W.A. Havenith, P.R. Taylor, C. Angeli, R. Cimiraglia, and K. Ruud, J. Chem. Phys. 120, 4619 (2004).
47 Y. Guo, K. Sivalingam, E.F. Valeev, and F. Neese, J. Chem. Phys. 144, 94111 (2016).
48

of the fully internally contracted NEVPT2 (FIC-NEVPT2). The DLPNO-NEVPT2 approach


can be combined with RIJK and RIJCOSX (Table 6), but the savings in computation time
are modest. Compared to the parenting FIC-NEVPT2 approach, the deviations are
negligible.
!def2-SVP def2/JK moread DLPNO-NEVPT2 PAL8
%moinp “cas2.gbw” # converged CASSCF(2,2)
%maxcore 3000 #more memory
# CAS 2 electrons in 2 orbitals
%casscf nel 2
norb 2
mult 3,1
end
* xyzfile 2 3 cu2.xyz

The RI approximation is also implemented for MRCI calculations. The results are
reported in Table 6. As a reminder, all integrals must be kept in memory. The smaller set
of RI integrals allows the computation of larger molecules. However, for molecules, that
can be treated exactly, the RI approximation does not accelerate the calculation but
rather slows it down significantly!
# def2/jk aux-basis recommended for CASSCF,NEVPT2 and MRCI
# def2-svp/c smaller aux-basis that can be sufficient for orca_mrci

! def2-SVP def2/JK MOREAD PAL4 noiter allowrhf nopop


%moinp "cas2.gbw" #converged CASSCF(2,2) orbitals
%maxcore 6000
%mrci citype mrddci3

intmode ritrafo

TSel 1e-8
newblock 1 *
refs cas(2,2) end
end
newblock 3 *
refs cas(2,2) end
end
davidsonopt none
end

* xyzfile 2 3 cu2.xyz

8.4 CASSCF(18,10) and NEVPT2


Quite often for metal dimers, the entire 3d manifold is selected in the active space, which
in the case of Cu2(OH)2 leads to CAS(18,10). The doubly non-magnetic orbitals are
exactly doubly occupied in the lowest singlet and triplet roots. This will ultimately lead
to convergence problems i.e. orbitals being rotated out of the active space. Although the
energy change is fairly small, and extended active space improves the NEVPT2 results.
We demonstrate this by reading the converged CASSCF(2,2) calculation and extending
the active space. The other 3d-orbitals are buried in the doubly occupied space of the
CAS(2,2) and partially mixed with the oxygen bridging ligands (5-20%).
54 55 56 57 58 59
-1.01235 -0.97939 -0.95283 -0.94870 -0.94574 -0.94345
2.00000 2.00000 2.00000 2.00000 2.00000 2.00000
-------- -------- -------- -------- -------- --------

0 Cu dz2 28.1 29.9 0.0 7.9 0.4 0.1


0 Cu dyz 0.0 0.0 44.6 0.0 0.0 0.0
0 Cu dx2y2 3.1 16.0 0.0 15.6 0.1 34.6
0 Cu dxy 0.1 0.2 0.0 1.9 45.0 0.1
1 Cu s 0.2 0.1 0.0 0.6 0.0 0.3
1 Cu pz 0.7 0.0 0.0 0.1 0.0 0.0
1 Cu dz2 28.1 29.9 0.0 7.9 0.4 0.1
1 Cu dxz 0.0 0.0 0.1 0.0 0.0 0.0
49

1 Cu dyz 0.0 0.0 44.6 0.0 0.0 0.0


1 Cu dx2y2 3.1 16.0 0.0 15.6 0.1 34.6
1 Cu dxy 0.1 0.2 0.0 1.9 45.0 0.1
2 O pz 0.0 0.5 0.1 2.8 0.1 0.0
2 O px 11.3 0.0 0.0 0.0 0.0 4.8
2 O py 0.0 0.0 2.0 0.0 1.1 0.0
3 O pz 0.0 0.5 0.1 2.8 0.1 0.0
3 O px 11.3 0.0 0.0 0.0 0.0 4.8
3 O py 0.0 0.0 2.0 0.0 1.1 0.0

In the input below, we rotate orbitals into the active space (94-103). For orbitals
optimization with orbitals, that are uncorrelated (2.0 or 0.0 occupation), the SuperCI
and DIIS overestimate the step-size. As a result, active orbitals can be rotated out of the
active space. Here, the best choice is the default converger (SuperCI_PT). It allows the
elimination of rotations with orbitals close to the critical occupation. The exact threshold
can be adjusted with the keyword “DThresh” (default=1e-6).
! def2-SVP RIJK conv def2-svp/jk moread PAL8 RI-NEVPT2
%maxcore 3000 #more memory
%moinp “cas2.gbw” # converged CASSCF(2,2) orbitals
%scf rotate
{54,101,90}{55,100,90}{56,99,90}{57,98,90}
{58,97,90}{59,96,90}{60,95,90}{61,94,90}
end
end
%casscf nel 18
norb 10
mult 3,1
end

* xyzfile 2 3 cu2.xyz

The program starts with a low gradient of ||g||= 0.0015, which emphasizes the similarity
between initial and final gradient. For many purposes the orbital optimization is not
necessary and a CAS-CI calculation with the extended active space will do it.
Alternatively, state-averaging, as suggested in Section 15.4, can help to avoid
convergence problems. As expected, the J-values have not changed (J=-56.7 cm-1).

Extension of the active space affects NEVPT2 and MRCI results. The improvements
reported in Table 7 are modest for NEVPT2. Further extension of the active space, e.g.
the bridging ligands, will get us closer to the experiment.

Table 7. NEVPT2 and DDCI3 J-couplings with extended active space (all 3d orbitals).

CAS(2,2) CAS(18,10)
NEVPT2 -117.4 -127.8
DDCI3 -435.9 -554.6

Note that uncontracted MRCI calculations with extended active space are time-
consuming and resource-demanding. Internally contracted approaches such as NEVPT2
are less restrictive on the size of the active space. Moreover, internally contracted
approaches can tackle large molecules, since the integrals are not kept in memory. The
fully internally contracted MRCI (FIC-MRCI) recently implemented in ORCA might be an
alternative in future applications.48 The current implementation of the FIC-MRCI does
not support parallel runs is thus not reported here.

48 K. Sivalingam, M. Krupicka, A.A. Auer, and F. Neese, J. Chem. Phys. 145, 54104 (2016).
50

8.5 Wave function Printing


ORCA by default prints the wave function in terms of configurations and their weights
(sum of squared CI coefficients), which might not be sufficient for all purposes. The wave
function can be printed in the basis of determinants. In this short section the necessary
keywords are described for the ORCA modules CASSCF, ICE and MRCI.

8.5.1 CASSCF and ICE


The wave function for the CASSCF/CASCI calculation can be printed by using the
“PrintWF” keyword as shown below.
! def2-SVP RIJK conv def2/jk moread PAL8 RI-NEVPT2
%maxcore 3000 #more memory
%moinp “cas2.gbw” # converged CASSCF(2,2) orbitals
%scf rotate
{54,101,90}{55,100,90}{56,99,90}{57,98,90}
{58,97,90}{59,96,90}{60,95,90}{61,94,90}
end
end
%casscf nel 18
norb 10
mult 3,1
printwf det # printing of the wave function in determinants
# in addition to the usual printing.
ci
TPrintWF 0.1 # cutoff for printing
end
end

* xyzfile 2 3 cu2.xyz

The “PrintWF” keyword takes two arguments “csf” and “det”. The “csf” option will print
the wave function in the usual configuration state function basis whereas the “det”
keyword will print the wave function in the determinant basis. Note that the printing of
the CSFs is only meaningful for high-spin states as the addressing of the CSFs is not
explained. The “TPrintWF” keyword in the CI section can be used to set the weight for
the printing of the configurations of the wave function. This option can also be used
while choosing different methods for the resolution of the CI problem.

8.5.2 MRCI
The MRCI block now also has the capacity for printing the CI wave function in the basis
of determinants. The wave function printing can be switched on by using the “PrintWF”
keyword in the MRCI block as shown below.
# def2/jk aux-basis recommended for CASSCF,NEVPT2 and MRCI
# def2-svp/c smaller aux-basis that can be sufficient for orca_mrci

! def2-SVP def2/jk MOREAD PAL4 noiter allowrhf nopop


%moinp "cas2.gbw" #converged CASSCF(2,2) orbitals
%maxcore 6000
%mrci citype mrddci3
tsel 1e-8 # tighter thresh
intmode ritrafo
printwf det # wave function printing in determinants
tprintwf 0.1 # cutoff for the printing
newblock 1 *
refs cas(2,2) end
end
newblock 3 *
refs cas(2,2) end
end
davidsonopt none
end

* xyzfile 2 3 cu2.xyz
51

Similar to the CASSCF section, the wave function is switched on by specifying the
“PrintWF det”. The threshold for the printing of the configurations is controlled by the
“TPrintWF” keyword. Only configurations with a weight greater than that specified by
TPrintWF are printed. The output for the DDCI3 calculations for the singlet state is
shown below in the presence of the determinant based printing option.
----------
CI-RESULTS
----------

The threshold for printing is 0.3 percent


The weights of configurations will be printed. The weights are summed over
all CSF's that belong to a given configuration before printing

STATE 0: Energy= -4118.424966772 Eh 0.00 eV 0.0 cm**-1

0.5341 : CFG h---h--- [20]


[20] 0.730817758

0.3998 : CFG h---h--- [02]


[02] -0.632315212

0.0017 : CFG h---h 78 [21]


[d2u] -0.028965827
[u2d] 0.028965827

0.0041 : CFG h---h 82 [21]


[d2u] -0.045399934
[u2d] 0.045399934

0.0087 : CFG h---h 84 [21]


[d2u] 0.066069489
[u2d] -0.066069489

0.0053 : CFG h---h100 [21]


[d2u] 0.051704571
[u2d] -0.051704571

0.0011 : CFG h---h 73 [12]


[du2] -0.023569365
[ud2] 0.023569365

The wave function coefficients are now shown in the basis of spin determinants. All the
determinants resulting from a given CFG are printed below each configuration
regardless of the coefficient. The spin-up electrons are shown by a "u" symbol and the
spin-down electrons are shown by a "d" symbol while the doubly occupied and
unoccupied orbitals are represented by the symbols 2 and 0 respectively. A convention
is chosen such that the singlet is written as shown below:
[ ]−[ ]
√2
The subscripts represent the electron labels and "u" and "d" represent the spin.
Therefore, in the single wave function shown above the [ ] and [ ] determinants
have opposite sign. Using the above convention and the coefficient of the wave function
in the basis of determinants, one can extract all possible information resulting from the
MRCI calculation.
52

9 A comment about CASSCF calculations on heavier elements


(lanthanide- and actinide-based systems)
In this section, we review some specific aspects of CASSCF calculations for lanthanide-
and actinide-based systems. Even though preceding comments on CASSCF calculations
for transition metal compounds also apply here, there are some pitfalls that are more
likely to be encountered in f-elements compounds or heavier elements in general.

9.1 Basis, RI-Basis and ECP


There is limited number of basis sets and auxiliary basis sets available for heavy
elements. Please ensure that the basis or auxiliary basis sets exists. Note that in older
versions of ORCA, the program did not necessarily abort in such a case!

All electron basis sets should take into account relativistic effects (DKH or ZORA). The
chosen basis sets should be “re-contracted” for the given scalar relativistic correction. In
our experience with lanthanides the SARC2 basis set is a good choice.49 These basis
sets are all electron basis sets designed for DFT, CASSCF and NEVPT2 calculations
including magnetic properties. They are also attractive, because there are pre-defined
auxiliary basis sets (/JK) available.

Other good choices are the Sapporo basis set or the generally contracted ANO-RCC basis
set. The AutoAux feature is useful, when you not sure which auxiliary basis to
pick.50 It is large and designed to mimic /JK and /C auxiliary basis sets. Note that the
AutoAux is fully decontracted.

Some more general remarks:


· Basis set handling and naming has changed in ORCA 4.0 (see the manual for more
details and trouble-shooting). For heavier elements ECPs are turned on by
default.
· Default frozen core settings have changed in ORCA 4.0 (check the manual for
more information on core correlation).
· Check for linear dependencies in the auxiliary basis (smallest eigenvalue < 10-7,
which might cause convergence problems in the CASSCF procedure. In case of
AutoAux, the basis can be further reduced (see manual). The smallest eigenvalues
of the RI metric are printed right before the CASSCF iterations. In the next version
linear dependencies will be automatically removed.
The smallest 3 eigenvalues: 7.997e-07 8.479e-07 1.700e-06

------------------
CAS-SCF ITERATIONS
------------------
...

9.2 Initial guess


For first row transition metal complexes, PAtom is a good choice as the resulting metal-
orbitals are ionic and at the frontier. For heavier elements, PAtom is not available. Thus,
preparing guess orbitals is mandatory.

49 D. Aravena, F. Neese, and D.A. Pantazis, J. Chem. Theory Comput. 12, 1148 (2016).
50 G.L. Stoychev, A.A. Auer, and F. Neese, J. Chem. Theory Comput. 13, 554 (2017).
53

Relativistic effects are important for these systems and should also be accounted to
when generating the guess. For a CAS space that consists of f-orbitals, good guess
orbitals are very ionic. The converged CASSCF orbitals have typically more than 80%
metal character. Aside from faster convergence, starting with ionic orbitals will help to
protect your active space during the orbital optimization.

The best guess is the fragment guess presented in Section 10. The only disadvantage
is in the number of preparation steps, that you to carry out. Nevertheless, this is the
recommended procedure. The alternatives discussed below are by far less efficient. In
the future the procedure might be further simplified.

DFT orbitals as well as “PModel” orbitals (default) tend towards more covalent f-
orbitals, but the can do the job as well. Note that PModel orbitals are essentially DFT
orbitals in the first iteration. Typically the first few DFT or HF iterations are quite
dramatic. Hence, converged DFT/HF orbitals are a better choice than PModel. Or in
other words, given the large number of core/internal to external rotations, it is a good
idea to start with orbitals that are pre-optimized for some of these rotations.

There many ways to generate orbitals. Keep in mind that these are just starting orbitals
and therefore overall convergence of the guess-calculation is not important. In case
of DFT or HF guess-calculations, the f-orbitals are not frontier orbitals and consequently
will not have the desired occupation. Below are a few ideas for start orbitals:

· The simplest choice is to start CASSCF with PModel (default). Inspect and rotate
the orbitals after the first CASSCF iteration as described in Section 2.2. Depending
on the system the active orbitals might be rotated out of the active space. Then
re-rotating and re-starting will get the calculation to convergence.

· UKS (alpha) or QRO orbitals from a simple DFT calculation.

· The DFT/HF orbitals of the ionized species can be better than PModel. For
example, to compute an U(III)-complex with 3 electrons in the f-orbitals
[CAS(3,7)] start with the RHF orbitals from U(VI) with 0 f-electrons.

Irrespective of the guess, the orbitals will need to be rotated! The initial gradient
will be large (compared to TM-Complexes), but the gradient will drop very rapidly
within the first few iterations.

9.3 Additional features concerning CASSCF calculations on f-elements systems

It is important to set a number of roots which includes complete spectroscopic terms


(2S+1L) in the CASSCF calculation. For instance, for calculations including only the
higher multiplicity roots of Ln(III) / An(III) complexes, you should consider the
following number of roots:

CeIII PrIII NdIII PmIII SmIII EuIII GdIII TbIII DyIII HoIII ErIII TmIII YbIII
ThIII PaIII UIII NpIII PuIII AmIII CmIII BkIII CfIII EsIII FmIII MdIII NoIII
54

7 21 35 35 21 7 1 7 21 35 35 21 7

You can certainly set a lower number of roots, if you want to focus only on the lowest
multiplets. For instance, for the Dy(III) free ion, if you are only interested in the spin-
orbit states stemming from the 6H term then you could decide to compute the 11 lowest
sextet roots only. However, this can have an impact on state mixing due to spin-orbit
coupling (SOC) in a quasi-degenerate perturbation theory (QDPT) calculation. Some care
and experimentation is recommended. These recommendations also apply if lower
multiplicity roots are required.

Scalar and spin-orbit relativistic effects are also very important in the description of
spectroscopic and magnetic properties of lanthanide- and actinide-based systems. Spin-
orbit effects can be described by the QDPT approach. In the context of molecular
magnetism, the approach has recently been reviewed.51 Moreover, the “picturechange”
option must be considered for property calculations.

9.4 Example: Investigation of the spectroscopic and magnetic properties of the


Cs2NaDyCl6 elpasolite

For a concrete example on the application of CASSCF calculations for real lanthanide
systems, we consider the spectroscopic and magnetic properties of the Cs2NaDyCl6
elpasolite. In this crystal, the Dy(III) center is surrounded by six chloride ions in an
octahedral arrangement. To keep the simplicity of the example, we just consider the
[DyCl6]3- fragment and ignore environment effects, although more realistic models for
this system can also be constructed.52

In this example, the SARC2 basis set is used. We start with preparing an initial guess. The
geometry, denoted as “dycl6.xyz”, is aligned so that the xyz axis points towards the
ligands. For this demonstration, we choose the UKS orbitals as guess. The compound has
a f9-configuration with the high-spin sextet state as the ground state.
!DKH DKH-DEF2-TZVP KDIIS xyzfile BP
%maxcore 2000
%basis
newgto Dy “SARC2-DKH-QZVP” end
end

*xyz -3 6
Dy 0.000000000 0.000000000 0.000000000
Cl -2.72 0.0 0.0
Cl 2.72 0.0 0.0
Cl 0.0 -2.72 0.0
Cl 0.0 2.72 0.0
Cl 0.0 0.0 -2.72
Cl 0.0 0.0 2.72
*

We explicitly used KDIIS in the DFT calculation to smoothen out convergence. Passing
the large initial fluctuations, the calculation signals convergence.
0 -14915.5437465821 0.000000000000 0.16409895 0.00115329 6.6899961 1.406279805

51 M. Atanasov, D. Aravena, E. Suturina, E. Bill, D. Maganas, and F. Neese, Coordination Chemistry Reviews
289–290, 177 (2015).
52 D. Aravena, M. Atanasov, F. Neese, Inorg. Chem., 2016, 55, 4457-4469
55

1 -14918.1213035206 -2.577556938537 0.27173413 0.00191386 6.0219582 1.249599488


2 -14921.9155834721 -3.794279951484 0.35127048 0.00248400 4.8199742 0.974355462
3 -14924.7990587092 -2.883475237137 0.19431191 0.00229227 2.8974524 0.565825432
4 -14923.8692218923 0.929836816918 0.31835271 0.00314092 0.5408713 0.379197897
5 -14925.0472495276 -1.178027635253 0.16128797 0.00128329 0.2073767 0.127890800
6 -14925.2879368394 -0.240687311876 0.08242954 0.00116018 0.1938458 0.134468471
7 -14925.6158737924 -0.327936952930 0.28792090 0.00212731 0.1324176 0.0899310650
...
***RMSP convergence achieved***
78 -14928.1546088356 0.000000683858 0.00023633 0.00000097 0.0000261 0.000014892
79 -14928.1546092732 -0.000000437556 0.00032045 0.00000126 0.0000261 0.000015916

We inspect the Loewdin orbital composition to identify the active orbitals. More
specifically we inspect the spin-up set.
------------------------------------------
LOEWDIN REDUCED ORBITAL POPULATIONS PER MO
-------------------------------------------
THRESHOLD FOR PRINTING IS 0.1%
SPIN UP
...
63 64 65
-.06196 0.06207 0.06536
.00000 1.00000 1.00000
------- -------- --------
0 Dy f0 65.0 0.4 0.8
0 Dy f+1 4.6 22.3 30.2
0 Dy f-1 1.9 24.1 36.6
0 Dy f+2 0.0 2.0 0.8
0 Dy f+3 15.8 21.4 12.8
0 Dy f-3 10.0 27.1 18.1
...

Orbitals 63-67 and 70 are occupied and strongly metal based f-orbitals. For comparison,
the converged CASSCF orbitals are pure f-orbitals (99% metal-based). Next we start the
actual CASSCF calculation reading the UKS guess orbitals. The orbitals need to be rotated
in order to fit the active space (81-87). When using UKS orbitals in CASSCF calculations,
the program automatically uses the MO coefficients of the spin-up set.
!DKH DKH-DEF2-TZVP xyzfile moread tightscf
%moinp “rhf.gbw” # guess orbitals
%maxcore 2000
%basis
newgto Dy “SARC2-DKH-QZVP” end
end
%scf rotate {63,81,90}{64,82,90}{65,83,90}{...} end
%casscf
nel 9
norb 7
nroots 21
mult 6
etol 1e-7 # reset energy convergence (overwritten by tightscf)
end
* xyzfile -3 6 dycl6.xyz

The calculations start with a large gradient of ||g|=65.7 and converges smoothly in 10
iterations. Note that the same calculation with PModel starts with a smaller gradient
(~20), but the calculation takes many more iterations to achieve convergence!
|g|| = 65.570594976 Max(G)= -35.228212477 Rot=394,0
|g|| = 0.976848225 Max(G)= -0.168279258 Rot=366,77
|g|| = 1.314439882 Max(G)= -0.207376148 Rot=282,52
|g|| = 0.241361301 Max(G)= 0.041085947 Rot=313,50
|g|| = 0.093756116 Max(G)= 0.014969442 Rot=314,51
|g|| = 0.011959253 Max(G)= 0.001871513 Rot=345,22
|g|| = 0.005455438 Max(G)= -0.000872837 Rot=317,49
|g|| = 0.001030188 Max(G)= -0.000151265 Rot=357,17
|g|| = 0.000066347 Max(G)= 0.000008914 Rot=82,54
|g|| = 0.000066366 Max(G)= -0.000007169 Rot=84,54
56

With the converged CASSCF orbitals, we compute the g-tensor (ground state),
magnetization and the susceptibility. Note that in the snippet below there are two “rel”
blocks. The global takes care of picture change effect and general settings regarding the
spin-orbit coupling operator, while the “rel” block in CASSCF sets the properties to be
computed. More refined options are documented in the ORCA manual. The ab initio
ligand field theory (AILFT) is available for CASSCF calculations with a single set of f-
orbitals. The analysis is enabled by adding “actorbs forbs” in the CASSCF block.
!DKH DKH-DEF2-TZVP moread tightscf
%moinp “cas.gbw” # converged CASSCF orbitals
%basis
newgto Dy “SARC2-DKH-QZVP” end
%rel
picturechange 2 # second order DKH SOC
fpFWtrafo false # recommended option for g-tensor
end
%casscf
nel 9
norb 7
nroots 21
mult 6
etol 1e-7 #reset energy convergence since we used tightscf

actorbs forbs # for ab initio ligand field analysis

rel
dosoc true
domagnetization true # to compute magnetization
dosusceptibility true# to compute magnetic susceptibility
gtensor true # to compute the g-tensor of
# the ground state
end
end
* xyzfile -3 6 dycl6.xyz

The converged calculation prints the SOC-corrected state.


Lowest eigenvalue of the SOC matrix: -14911.32612911 Eh
Energy stabilization: -4851.20493 cm-1
Eigenvalues: cm-1 eV Boltzmann populations at T = 300.000 K
0: 0.0000 0.0000 9.85e-02
1: 0.0000 0.0000 9.85e-02
2: 18.6261 0.0023 9.01e-02
3: 18.6261 0.0023 9.01e-02
4: 18.6514 0.0023 9.01e-02
5: 18.6514 0.0023 9.01e-02
6: 85.8167 0.0106 6.53e-02
7: 85.8167 0.0106 6.53e-02
8: 175.0115 0.0217 4.26e-02
9: 175.0115 0.0217 4.26e-02
10: 175.0192 0.0217 4.26e-02
11: 175.0192 0.0217 4.26e-02
12: 213.2292 0.0264 3.54e-02
13: 213.2292 0.0264 3.54e-02
14: 213.2343 0.0264 3.54e-02
15: 213.2343 0.0264 3.54e-02
16: 3024.5421 0.3750 4.94e-08
17: 3024.5421 0.3750 4.94e-08
18: 3025.5097 0.3751 4.92e-08
19: 3025.5097 0.3751 4.92e-08
20: 3025.5307 0.3751 4.92e-08
21: 3025.5307 0.3751 4.92e-08
22: 3075.8603 0.3814 3.86e-08
23: 3075.8603 0.3814 3.86e-08

As well as the magnetic properties:


-------------------------------------------------
SOC CORRECTED MAGNETIZATION AND/OR SUSCEPTIBILITY
-------------------------------------------------

Do magnetization? = True
57

Minimum Field (G) = 0.000


Maximum Field (G) = 70000.000
Number of field points = 15
Do susceptibility? = True
Minimum Temperature (K) = 1.000
Maximum Temperature (K) = 300.000
Number of temperatures = 300
Lebedev grid precision = 5
Npoints for differentiation = 5
Field step = 100.000

Calculating magnetization ...

-----------------------------------------------------------
FIELD DEPENDENT MAGNETIZATION
-----------------------------------------------------------
TEMPERATURE (K) M. FIELD (Gauss) MAGNETIZATION (B.M.)
-----------------------------------------------------------

4.00 0.00 0.000000


4.00 5000.00 1.541865
4.00 10000.00 2.897042
4.00 15000.00 3.978282
4.00 20000.00 4.798456
4.00 25000.00 5.414143
4.00 30000.00 5.882877
4.00 35000.00 6.248377
4.00 40000.00 6.540476
4.00 45000.00 6.778939
4.00 50000.00 6.976992
4.00 55000.00 7.143753
4.00 60000.00 7.285741
4.00 65000.00 7.407776
4.00 70000.00 7.513528
-----------------------------------------------------------

Calculating susceptibility ...

-----------------------------------------------------------
TEMPERATURE DEPENDENT MAGNETIC SUSCEPTIBILITY
-----------------------------------------------------------
STATIC FIELD (Gauss) TEMPERATURE (K) chi*T (cm3*K/mol)
-----------------------------------------------------------

0.00 1.00 4.885933


0.00 2.00 5.608100
0.00 3.00 6.330229
0.00 4.00 7.048052
0.00 5.00 7.747386
0.00 6.00 8.407303
0.00 7.00 9.009442
0.00 8.00 9.543478
0.00 9.00 10.007350
0.00 10.00 10.404896

In the present case, we can compare the computed magnetic susceptibility curve for the
DyCl63- fragment with the corresponding experimental data.53

53 Dunlap and Shenoy, 1975, PRB, 12, 7, 2716


58

Figure 18 Experimental (blue) and calculated (red) magnetic susceptibility curves.

The g-tensor for the ground spin-orbit state is also printed out with, at first, the raw g-
tensor, followed by the three main values of this tensor after diagonalization and the
associated main direction, in the molecular reference coordinate system (see Section
7.4).
-------------------
ELECTRONIC G-MATRIX
-------------------

g-matrix:
6.510988 1.352594 0.000001
1.344898 -6.548245 -0.000013
0.000002 -0.000011 6.652526

g-factors:
6.648437 6.652526 6.686480 iso = 6.662481

g-shifts:
4.646118 4.650206 4.684161 iso = 4.660162

Eigenvectors:
1.000000 0.000062 -0.000000
-0.000000 -0.000020 -1.000000
-0.000062 1.000000 -0.000020

Finally, the matrix elements of ligand field model Hamiltonian are printed out:
H(fm3 ,fm3 )= 0.054851750 a.u. = 1.493 eV = 12038.6 cm**-1
H(fm2 ,fm3 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(fm2 ,fm2 )= 0.053469747 a.u. = 1.455 eV = 11735.3 cm**-1
H(fm1 ,fm3 )= -0.000595617 a.u. = -0.016 eV = -130.7 cm**-1
H(fm1 ,fm2 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(fm1 ,fm1 )= 0.054545752 a.u. = 1.484 eV = 11971.4 cm**-1
H(f0 ,fm3 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f0 ,fm2 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(f0 ,fm1 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(f0 ,f0 )= 0.055313741 a.u. = 1.505 eV = 12140.0 cm**-1
H(f1 ,fm3 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f1 ,fm2 )= -0.000000001 a.u. = -0.000 eV = -0.0 cm**-1
H(f1 ,fm1 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f1 ,f0 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f1 ,f1 )= 0.054544252 a.u. = 1.484 eV = 11971.1 cm**-1
59

H(f2 ,fm3 )= 0.000000001 a.u. = 0.000 eV = 0.0 cm**-1


H(f2 ,fm2 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f2 ,fm1 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(f2 ,f0 )= 0.000000108 a.u. = 0.000 eV = 0.0 cm**-1
H(f2 ,f1 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f2 ,f2 )= 0.054083942 a.u. = 1.472 eV = 11870.1 cm**-1
H(f3 ,fm3 )= -0.000000000 a.u. = -0.000 eV = -0.0 cm**-1
H(f3 ,fm2 )= -0.000000001 a.u. = -0.000 eV = -0.0 cm**-1
H(f3 ,fm1 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(f3 ,f0 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(f3 ,f1 )= 0.000595127 a.u. = 0.016 eV = 130.6 cm**-1
H(f3 ,f2 )= 0.000000000 a.u. = 0.000 eV = 0.0 cm**-1
H(f3 ,f3 )= 0.054853298 a.u. = 1.493 eV = 12038.9 cm**-1

Comparison between the CASSCF energies and the same energies recomputed using the
newly determined ligand field parameters is printed too:
AI-Root 0: E(AI)= 0.000 eV -> LF-Root 1: 0.000 eV S= 1.000 Delta= -0.000 eV
AI-Root 1: E(AI)= 0.000 eV -> LF-Root 0: 0.000 eV S= 1.000 Delta= 0.000 eV
AI-Root 2: E(AI)= 0.000 eV -> LF-Root 2: 0.000 eV S= 1.000 Delta= -0.000 eV
AI-Root 3: E(AI)= 0.020 eV -> LF-Root 3: 0.019 eV S= 1.000 Delta= 0.000 eV
AI-Root 4: E(AI)= 0.020 eV -> LF-Root 4: 0.019 eV S= 0.998 Delta= 0.000 eV
AI-Root 5: E(AI)= 0.020 eV -> LF-Root 5: 0.019 eV S= 0.998 Delta= 0.000 eV
AI-Root 6: E(AI)= 0.032 eV -> LF-Root 6: 0.031 eV S= 0.994 Delta= 0.001 eV
AI-Root 7: E(AI)= 0.032 eV -> LF-Root 7: 0.031 eV S= 0.994 Delta= 0.001 eV
AI-Root 8: E(AI)= 0.038 eV -> LF-Root 9: 0.036 eV S= 1.000 Delta= 0.001 eV
AI-Root 9: E(AI)= 0.038 eV -> LF-Root 8: 0.036 eV S= 1.000 Delta= 0.001 eV
AI-Root 10: E(AI)= 0.038 eV -> LF-Root 10: 0.036 eV S= 1.000 Delta= 0.001 eV
AI-Root 11: E(AI)= 0.937 eV -> LF-Root 13: 0.939 eV S= 1.000 Delta= -0.002 eV
AI-Root 12: E(AI)= 0.937 eV -> LF-Root 11: 0.939 eV S= 1.000 Delta= -0.002 eV
AI-Root 13: E(AI)= 0.937 eV -> LF-Root 12: 0.939 eV S= 1.000 Delta= -0.002 eV
AI-Root 14: E(AI)= 0.951 eV -> LF-Root 14: 0.950 eV S= 1.000 Delta= 0.002 eV
AI-Root 15: E(AI)= 0.951 eV -> LF-Root 15: 0.950 eV S= 1.000 Delta= 0.002 eV
AI-Root 16: E(AI)= 0.951 eV -> LF-Root 16: 0.950 eV S= 1.000 Delta= 0.002 eV
AI-Root 17: E(AI)= 0.962 eV -> LF-Root 17: 0.955 eV S= 1.000 Delta= 0.007 eV
AI-Root 18: E(AI)= 4.337 eV -> LF-Root 18: 4.336 eV S= 1.000 Delta= 0.001 eV
AI-Root 19: E(AI)= 4.337 eV -> LF-Root 19: 4.336 eV S= 1.000 Delta= 0.001 eV
AI-Root 20: E(AI)= 4.337 eV -> LF-Root 20: 4.336 eV S= 1.000 Delta= 0.001 eV

Energies of the so-called CASSCF ligand field orbitals also appear in the output. As
explained in section 3, the ligand field orbitals can be visualized (Figure 19).

Figure 19. Ligand field orbitals and splitting parameters.

The ligand field one electron eigenfunctions:


Orbital Energy (eV) Energy(cm-1) f-3 f-2 f-1 f0 f+1 f+2 f+3
1 0.000 0.0 -0.000000 1.000000 -0.000000 -0.000000 0.000001 0.000000 -0.000000
2 0.017 134.8 -0.612852 -0.000000 -0.790174 -0.000000 -0.000007 0.006072 0.000005
60

3 0.017 134.8 0.000005 0.000001 0.000006 0.000000 -0.790985 -0.000181 0.611836


4 0.017 134.8 -0.003721 0.000000 -0.004798 0.000088 0.000143 -0.999982 -0.000110
5 0.050 404.7 0.000023 -0.000001 -0.000018 0.000008 0.611836 0.000000 0.790985
6 0.050 404.7 -0.790148 -0.000000 0.612832 -0.010164 0.000018 -0.000001 0.000023
7 0.050 404.7 0.008032 -0.000000 -0.006230 -0.999948 0.000005 -0.000088 0.000006

In the present case, due to the octahedral symmetry, it is possible to parametrize the
energy difference between the ligand field orbitals (Figure 19) using two parameters
from the Angular Overlap Model (AOM)

eρ < 1 / 2Χ2 , 3 / 10Χ1


eο < 2 / 5Χ1 .

In the present example, for the [DyCl6]3- fragment, Δ1 = 134.8 cm-1 and Δ2 = 404.7 cm-1,
thus eσ = 161.9 cm-1 and eπ = 53.9 cm-1. For further information concerning the
parametrization of the AOM (i.e. for lower symmetry and/or other ligands), see W.
Urland, Chem. Phys., 1976, 14, 393-401 and C. E. Schäffer, Struct. Bond., 1968 ,5, 67-95.
Finally, from the AILF analysis we also get access to the SOC parameter ξ and to the
inter-electronic repulsion parameters F2, F4 and F6 (unormalized Slater-Condon
parameters).
F2 = 0.500695983 a.u. = 13.625 eV = 109890.1 cm**-1

SOC constant zeta = 0.215 eV = 1737.8 cm**-1

Typically the Slater-Condon parameters are processed in their normalized form


F2 < F2 / 225,
F4 < F 4 / 1089,
F6 < F6 / 7361.64,

which can be further translated into the Racah parameters E1, E2 and E3 such that

E1 < ∋70F ∗ 231F ∗ 2002F (


2 4 6
/ 9,
E 2
< ∋F , 3F ∗ 7F ( / 9,
2 4 6

E3 < ∋5F ∗ 6F , 91F ( / 3.


2 4 6

The parameter E3 is responsible for the energy splitting between the terms of highest
multiplicity (Smax), while E1 and E2 are responsible for the energy splitting between the
terms of lower multiplicity than the highest (i.e. Smax-1, Smax-2, …). If only the highest
multiplicity roots are computed, like in the example, there is only one inter-electronic
repulsion parameter (F2), which is related to E3 such that E3 = F2 / 135.

10 Fragment Derived Guess (orca_mergefrag)


In this section, we discuss a less known feature/module in ORCA that is
orca_mergefrag. This feature can be used to generate ionic guess orbitals, which is
not easy for complexes with heavy elements or multiple metal centers. The idea is fairly
simple:
61

I) Divide the molecule into the fragments (metal and ligand).


II) Obtain molecular orbitals for each fragment. Naturally the metal fragment will
have purely metal based orbitals!
III) Merge the fragment orbitals to get the single set of MOs that is used as guess
for the actual calculation.

We sketch the process for the [DyCl6]3- complex studied in the previous section. First,
we prepare and run the ligand calculation denoted as “L”. Therefore, we delete the metal
from the geometry. For the UKS calculation, charge and multiplicity need to be adjusted
accordingly.
!DKH DKH-DEF2-TZVP KDIIS BP UKS
%maxcore 2000
%basis
newgto Dy “SARC2-DKH-QZVP” end
end

*xyz -6 1
Cl -2.72 0.0 0.0
Cl 2.72 0.0 0.0
Cl 0.0 -2.72 0.0
Cl 0.0 2.72 0.0
Cl 0.0 0.0 -2.72
Cl 0.0 0.0 2.72
*

The calculation converges rapidly and the resulting orbitals are stored in the “L.gbw”
file. Next, we run a CASSCF calculation on the Dy ion (metal fragment), where once again
charge and multiplicity are adjusted. The calculation is referred to as “CAS”. We use
rotated PModel orbitals as guess for the Dy(III) ion.
!DKH DKH-DEF2-TZVP printbasis printmos

%basis
newgto Dy "sarc2-dkh-qzvp" end
end

# rotate pmodel for f-orbitals to enter the active space


%scf rotate
{27,36,90}{28,37,90}{29,38,90}{30,39,90}{31,40,90}{32,41,90}{33,42,90} end
end
%pal nprocs 4
end

%casscf nel 9
norb 7
mult 6
nroots 21
end

*xyz 3 6
Dy 0.000000000 0.000000000 0.000000000
*

We refer the converged CASSCF orbitals as “CAS.gbw”. Now, it is time to merge the
fragment orbitals to get the final set of guess molecular orbitals. This is done from the
command line calling:
orca_mergefrag L.gbw CAS.gbw LCAS_merged.gbw

The program generates the desired LCAS_merged.gbw file, which can be used as guess
for the actual molecule. The sequence of arguments matters for orca_mergefrag. The
CASSCF orbitals should be second argument. In that case, the LCAS_merged.gbw is
62

generated for a merged structure, where the metal-fragment is the last element in the
xyz input block:
*xyz -3 6
Cl -2.72 0.0 0.0
Cl 2.72 0.0 0.0
Cl 0.0 -2.72 0.0
Cl 0.0 2.72 0.0
Cl 0.0 0.0 -2.72
Cl 0.0 0.0 2.72
Dy 0.000000000 0.000000000 0.000000000 # <- CAS fragment
*

It is a good idea to inspect and if necessary rotate the guess orbitals. As can be seen from
the gradient progression below, the CASSCF calculation on the full molecule starts with a
smaller gradient and no convergence problems are met.
||g|| = 7.395466558 Max(G)= -2.436430117 Rot=386,6
||g|| = 1.423366854 Max(G)= -0.303816590 Rot=104,70
||g|| = 0.254083028 Max(G)= 0.035885507 Rot=101,59
||g|| = 0.232270719 Max(G)= 0.037578604 Rot=314,51
||g|| = 0.136055209 Max(G)= 0.022725572 Rot=314,51
||g|| = 0.008657623 Max(G)= 0.001100328 Rot=314,51
||g|| = 0.003405623 Max(G)= -0.000415905 Rot=316,48
||g|| = 0.000392150 Max(G)= -0.000051813 Rot=304,23
||g|| = 0.000090863 Max(G)= 0.000012788 Rot=159,63
||g|| = 0.000090868 Max(G)= 0.000012781 Rot=159,63

11 Manipulation of the ORCA GBW File (orbitals)


This section illustrates how to edit the molecular orbital coefficients. The procedure is
for the expert user! As an example, we generate a “custom guess” for the [CrCl6]3-
complex studied in Section 15.1. The goal is to have the metal 3d and 4d orbitals in the
position of the active orbitals. The resulting .gbw file can then be used as a guess for the
CAS(3,10) calculation.

Starting point is an arbitrary .gbw file for the given system and basis e.g.
!SV def2/J BP printbasis normalprint
%base “bpguess” #defines the basename of the files created by ORCA
%output print[p_MOs] 1 end # prints the MO coefficients
* int -3 4
Cr 0 0 0 0.0000 0.000 0.000
Cl 1 0 0 2.4674 0.000 0.000
Cl 1 2 0 2.4674 90.000 0.000
Cl 1 2 3 2.4674 90.000 90.000
Cl 1 2 3 2.4674 90.000 180.000
Cl 1 2 3 2.4674 180.000 0.000
Cl 1 2 3 2.4674 90.000 270.000
*

These are UKS calculations with two orbitals sets. In the following, we work on the alpha
set, as the CASSCF program only reads the first set of orbitals (beta orbitals are
discarded). Inspection of the Loewdin population analysis shows that orbitals 63-67 are
already pre-dominantly metal 3d orbitals (60%). The orbitals 68-72 are not the desired
metal 4d orbitals. Hence, we want to modify the corresponding MO coefficients.
66 67 68 69 70 71
0.32474 0.32474 0.42957 0.58799 0.86784 0.86784
0.00000 0.00000 0.00000 0.00000 0.00000 0.00000
-------- -------- -------- -------- -------- --------
0 Cr s 0.0 0.0 89.7 64.0 0.0 0.0
0 Cr dz2 62.6 0.0 0.0 0.0 0.0 0.0
0 Cr dx2y2 0.0 62.6 0.0 0.0 0.0 0.0
1 Cl s 0.1 0.3 1.6 1.7 1.6 42.3
63

...

ORCA uses a binary file (.gbw, .qro, .uco, .uno and .nat) to store the MO coefficients and
related data. The input file above produces a bpguess.gbw file on disk. It is possible to
convert the file into human readable ASCII file with the orca_2mkl module. Call the
orca_2mkl program from the shell specifying the name of the .gbw file omitting the
“.gbw”:
orca_2mkl bpguess –mkl -u

The program produces a file named bpguess.mkl on disk. The .mkl file (Molekel format)
has several entries grouped into blocks starting with the $ symbol and a keyword-string.
Blocks are closed with “$END”. The alpha orbitals are stored after the flag
$COEFF_ALPHA.
$COEFF_ALPHA
a1g a1g a1g a1g a1g
-214.2072749 -100.1707337 -100.1706031 -100.1706031 -100.1705474
0.9956322 0.0000448 0.0000000 0.0000000 -0.0000000
0.0141653 -0.0000827 -0.0000000 -0.0000000 0.0000000
...
$END

For our purpose this is the only block of interest. The MOs are stored as column vectors
with a maximum of 5 columns per “batch”. Each MO starts with the irreducible
representation (a1g in our example), orbital energy followed by rows with the actual
coefficients. The MO coefficients are not normalized and stored in the following order:
loop atoms (as listed in the input geometry)
loop angular momentum L (e.g. s,p,d,...)
loop M_L (e.g. for L=1: px,py,pz)
store coefficient of AO(atom,L and M_L)

It is essentially the same ordering and printing ORCA uses in the output to print the MO
coefficients (!PrintMOs).
%output print[p_MOs] 1 end # prints the MO coefficients

Counting the number of columns we inspect orbitals 68-72 to be edited – below a


snippet of orbitals 65-69.
a1g a1g a1g a1g a1g
0.2560556 0.3247428 0.3247428 0.4295695 0.5879879
-0.0000000 -0.0000000 0.0000000 0.0011933 0.0432268
0.0000000 0.0000000 -0.0000000 -0.0012902 -0.1090271
-0.0000000 -0.0000000 0.0000000 0.0256604 0.6599043
0.0000000 0.0000000 -0.0000000 -1.1502698 -2.4864421
-0.0000000 0.0000000 0.0000000 2.2522996 0.9107544
-0.0000000 -0.0000000 -0.0000000 -0.0000000 -0.0000000
-0.0000000 0.0000000 -0.0000000 -0.0000000 -0.0000000
-0.0000000 -0.0000000 -0.0000000 -0.0000000 0.0000000
0.0000000 0.0000000 0.0000000 0.0000000 0.0000000
0.0000000 -0.0000000 0.0000000 0.0000000 0.0000000
0.0000000 0.0000000 0.0000000 0.0000000 -0.0000000
0.0000000 -0.7561386 -0.0019474 0.0000000 -0.0000000
0.0096330 0.0000000 -0.0000000 0.0000000 0.0000000
-0.0192398 -0.0000000 0.0000000 0.0000000 0.0000000
-0.0000000 0.0019474 -0.7561386 0.0000000 0.0000000
0.7256976 -0.0000000 -0.0000000 0.0000000 0.0000000
0.0000000 -0.2429614 -0.0006257 0.0000000 -0.0000000
0.0027890 -0.0000000 0.0000000 -0.0000000 -0.0000000
-0.0055703 0.0000000 -0.0000000 -0.0000000 -0.0000000
64

-0.0000000 0.0006257 -0.2429614 0.0000000 -0.0000000


0.2101039 0.0000000 0.0000000 0.0000000 -0.0000000

For the SV basis, Cr has 5 contracted s-functions, 2 sets of contracted p-function and 2
sets of contracted d-functions. Since Cr is the first atom in the list, the first 11
coefficients, marked green, are the Cr s and p functions. The next 5 are first shell of metal
d, followed by the second shell of more diffuse metal d-orbitals. The MOs 63-67 are
mostly the first shell. The neighboring orbitals have no contribution from any d orbitals.
In this example, we set a significant weight on the second shell of d-orbitals (MOs 68-
72). Reading the orbitals, ORCA will do Gram-Schmidt orthogonalization that generates
beautiful atom-like 3d and 4d orbitals.
a1g a1g a1g a1g a1g
0.2560556 0.3247428 0.3247428 0.4295695 0.5879879
-0.0000000 -0.0000000 0.0000000 0.0011933 0.0432268
0.0000000 0.0000000 -0.0000000 -0.0012902 -0.1090271
-0.0000000 -0.0000000 0.0000000 0.0256604 0.6599043
0.0000000 0.0000000 -0.0000000 -1.1502698 -2.4864421
-0.0000000 0.0000000 0.0000000 2.2522996 0.9107544
-0.0000000 -0.0000000 -0.0000000 -0.0000000 -0.0000000
-0.0000000 0.0000000 -0.0000000 -0.0000000 -0.0000000
-0.0000000 -0.0000000 -0.0000000 -0.0000000 0.0000000
0.0000000 0.0000000 0.0000000 0.0000000 0.0000000
0.0000000 -0.0000000 0.0000000 0.0000000 0.0000000
0.0000000 0.0000000 0.0000000 0.0000000 -0.0000000
0.0000000 -0.7561386 -0.0019474 0.0000000 -0.0000000
0.0096330 0.0000000 -0.0000000 0.0000000 0.0000000
-0.0192398 -0.0000000 0.0000000 0.0000000 0.0000000
-0.0000000 0.0019474 -0.7561386 0.0000000 0.0000000
0.7256976 -0.0000000 -0.0000000 0.0000000 0.0000000
0.0000000 -0.2429614 -0.0006257 100.0000000 -0.0000000
0.0027890 -0.0000000 0.0000000 -0.0000000 -100.0000000
-0.0055703 0.0000000 -0.0000000 -0.0000000 -0.0000000
-0.0000000 0.0006257 -0.2429614 0.0000000 -0.0000000
0.2101039 0.0000000 0.0000000 0.0000000 -0.0000000

Having edited the .mkl file, the file is converted back to the regular .gbw format using
orca_2mkl

orca_2mkl bpguess -gbw

Note that the orbitals in the resulting bpguess.gbw file are not normalized and non-
orthogonal. However, ORCA takes care of it while reading the .gbw file.
#Checking the input orbitals with noiter
!sv bp printbasis normalprint noiter
!moread
%moinp “bpguess.gbw”
* int -3 4

The reduced Loewdin analysis now correctly prints orbitals 68-72 as pre-dominantly
metal d-orbitals. The edited bpguess.gbw file is ready to be used
66 67 68 69 70 71
0.32474 0.32474 0.42957 0.58799 0.86784 0.86784
0.00000 0.00000 0.00000 0.00000 0.00000 0.00000
-------- -------- -------- -------- -------- --------
0 Cr dz2 62.6 0.0 83.7 0.0 0.0 0.0
0 Cr dxz 0.0 0.0 0.0 95.8 0.0 0.0
0 Cr dyz 0.0 0.0 0.0 0.0 95.8 0.0
0 Cr dx2y2 0.0 62.6 0.0 0.0 0.0 83.8

Editing the MKL file is best done by a small script.


65

12 p-Phenylenediamine (PPD) ground state - organic molecule


For completeness, we have added an example for an organic molecule – C6H4(NH2)2. This
molecule has delocalized electrons belonging to the benzene ring, which should be
described by a CAS(6,6) active space consisting of the three and three * orbitals. We
illustrate the selection of orbitals for such systems. Furthermore, we demonstrate how
to use the orca_loc program to identify lone-pairs.

The structure is aligned so that the molecule essentially lies in the x,y plane of the
coordinate system. This simplifies the identification of the -system depicted in Figure
17. In our experience, MP2 natural orbitals are a good starting point for the
identification of the active space. We use the RI approximation to speed up the
calculation. Natural orbitals from the unrelaxed density matrix are cheap and equally
good as starting orbitals (compared to natural orbitals from relaxed densities).
# RHF = restricted Hatree-Fock calculation
# DEF2-TZVP = triple-zeta basis set
# DEF2/J in conjunction with RIJCOSX to approximate the Fock operator
# DEF2-TZVP/C for the RI approximation in the MP2 section
# BOHRS = the unit of input geometry is in Bohr not in Angström
#
! RHF def2-TZVP BOHRS def2-TZVP/J RIJCOSX def2-TZVP/C RI-MP2
%maxcore 600
%mp2 density unrelaxed # cheaper mp2 densities
natorbs true # generates natural orbital file .mp2nat
end

# This geometry is taken from the supporting information of the paper:


# R. Send, M. Kühn, and F. Furche, J. Chem. Theory Comput. 7(8), 2376 (2011).
#
* xyz 0 1
c -2.25866526908469 1.31069703573197 -0.03890132740005
c -2.25862894906931 -1.31072721575442 0.05943295935235
c 0.00004113616061 -2.67501441686706 0.10690773212946
c 2.25866487869290 -1.31066640155145 0.05957021971887
c 2.25859837611157 1.31076900588401 -0.03878216913493
c -0.00005341198088 2.67507600749642 -0.08595384918212
n -0.00005616581783 5.32257534102916 -0.31225229131962
n 0.00000654177395 -5.32246493724021 0.33343657553687
h -4.04641808709996 2.31200286749062 -0.08850127780381
h -4.04634920958691 -2.31210657869357 0.10866632361833
h 4.04641683322748 -2.31198264095411 0.10894712212357
h 4.04632778359826 2.31212919613840 -0.08825146488559
h 1.56842673349648 6.16356336525531 0.37349462735647
h -1.56848011566668 6.16366447966973 0.37350714299725
h -1.56878868304699 -6.16347745476038 -0.35149324604992
h 1.56824237532426 -6.16377238160470 -0.35241434795340
*

The input above generates MP2 natural orbitals, where the -orbitals are identified at
the HOMO-LUMO gap. The program stores the orbitals as .mp2nat file, which can be used
like any other GBW file. Next we start the actual CASSCF calculation, where the geometry
is abbreviated as “PPD.xyz”.
#
# CASSCF(6,6) calculation for PPD
#
%MaxCore 900
!def2-TZVP BOHRS moread
%moinp "PPD.mp2nat" #mp2 natural orbitals with pi orbitals in place
%casscf
nel 6
norb 6
mult 1
end
* xyzfile 0 1 PPD.xyz
66

The output shows the occupation numbers of the active orbitals as well as their
separation from the inactive and virtual orbital subspaces.
--- Energy gap subspaces: Ext-Act = -0.352 Act-Int = 0.012
--- current l-shift: Up(Ext-Act) = 1.95 Dn(Act-Int) = 1.59
N(occ)= 1.97459 1.93246 1.93072 0.07552 0.06477 0.02194
||g|| = 0.104038220 Max(G)= -0.030815266 Rot=47,27

The -orbitals are well correlated as can be seen from the printing above. The energy
gap between doubly occupied and active orbitals is fairly small (0.012 Eh), while the
virtual and active orbitals are overlapping in energy. Note that this can be a problem for
the convergers SuperCI/DIIS and NR, where a suitable level-shift is needed (on by
default). The SuperCI_PT (program default) does not require any level shifts and finishes
in 8 iterations. It is good practice, to inspect the active orbitals after convergence. The
best way to do it is to visualize the orbitals (Figure 20).

# 26
# 27

# 28 # 29

# 31
# 30

Figure 20. Depicted the converged CASSCF(6,6) orbitals of the PPD molecule.

Depending on the chemical processes or properties studied, the lone pairs on the
nitrogen atoms may also be involved. In that case, one should include these orbitals into
the active space. The converged CASSCF(6,6) wave function is a good starting point to
add lone pairs. However, the lone pairs are in the doubly occupied space and heavily
mixed with the other orbitals of the same subspace at least for canonical orbitals.
However, lone pairs are easily identified if localized orbitals are used.54

54 keywords “intorbs locorbs” in the %casscf block


67

Here, we show how to get localized orbitals with the orca_loc program that is directly
called from the shell. The program explains the necessary input structure if it is called
without an input file.

The inactive space ranges from orbital 0 to 25. Below is the necessary input file, where a
major part of the input parameters are essentially the default settings repeated. The
important parts are the first 6 lines (input GBW, output GBW, operator, and orbital
range).
PPD.gbw # input orbitals
LOC.gbw # output orbitals
0 # operator:alpha here (CASSCF), 1 for beta
0 # orbital window: first orbital to be localized e.g. first active
25 # orbital window: last orbital to be localized e.g. last active
50 # maximum number of iterations
1 # localization method: 1=PIPEK-MEZEY

With the input at hand, we call the localization program from the command line.
orca_loc localization.input

The localization program prints a summary, where the lone pairs should appear as
“strongly localized” orbitals.
...
FOUND - 10 strongly local MO`s
- 16 two center bond MO`s
- 0 significantly delocalized MO`s

Rather strongly localized orbitals:


MO 9: 7N - 1.000811
MO 8: 7N - 0.955801
MO 7: 6N - 0.955827
MO 6: 6N - 1.000811

We visually inspect orbitals 6-9 and identify the lone pairs depicted in Figure 21.

#7

#8

Figure 21. Lone pairs identified from the localization procedure.


68

To prepare the CASSCF calculation with extended active space, the lone pairs need to be
placed as highest doubly occupied orbitals (in this case position 24 and 25), next to the
active space.
!DEF2-TZVP BOHRS
!moread
%MaxCore 900
%moinp "LOC.gbw" #CAS(6,6) with localized inactives

# Rotate lone pair MOs into the active space.


%scf
rotate
{24,7,90}
{25,8,90}
end
end

%casscf
nel 10
norb 8
mult 1
end
* xyzfile 0 1 PPD.xyz

Since the lone pairs are barely correlated to orbitals in the ground state calculation,
their occupation numbers are very close to 2.0 as can be seen from the output.
--- Energy gap subspaces: Ext-Act = 0.015 Act-Int = 0.077
N(occ)= 1.99938 1.99886 1.96313 1.90729 1.90480 0.09882 0.09271 0.03501
||g|| = 0.011086382 Max(G)= -0.007978892 Rot=30,1

Orbitals with and without lone pairs are very similar (small initial gradient). For the
ground state calculations, the lone pairs should not be included in the orbitals
optimization to avoid convergence problems. If you want to run the calculation
nevertheless, we recommend to use the default settings (SuperCI_PT) or the NR methods
with a sufficient level-shift. In this example both methods are equally efficient (6-10
iterations).

13 Adenine spectra – using symmetry


This section shows the symmetry handling in ORCA, which is exemplified on the vertical
excitation spectra of the adenine molecule. Symmetry handling is enabled with the
keyword “USESYM”. The molecule does not need to be aligned in specific way. ORCA
automatically detects point group symmetry55 and rotates the molecules accordingly. A
note of warning, geometry optimizations and scans with the activated symmetry feature
will not respect the point group and are not recommended!56 Although symmetry in
ORCA is not fully exploited, for multireference calculations substantial time-savings can
be achieved. Furthermore, it allows to directly target specific excited states, which
otherwise would require the inclusion of many more roots.

The adenine molecule has Cs point group symmetry, so there are only two irreducible
representations (A’, A’’). The states of interest are the lowest → ∗ and → ∗ excited
states reported in a benchmark study of Schapiro et al.57 To simplify the calculation we

55 keyword: “method symthresh 1e-2 end” (symmetry detection will slightly modify the geometry!)
56 Unrelaxed scans can be emulated with the multixyz-file option
57 I. Schapiro, K. Sivalingam, and F. Neese, J. Chem. Theory Comput. 9, 3567 (2013)
69

opt for state-averaged orbitals. Geometry is taken from the original article (MP2
optimized).

We follow the recipe of section 12 and start with the → ∗ excitation using the state-
averaged CASSCF methodology. Guess orbitals are once again generated from RI-MP2
natural orbitals. When using symmetry it is important to use the keywords USESYM
throughout all calculations including the guess!
# USESYM enables symmetry handling
# def2/JK RIJK for the RI approximation in the HF part of the calculation
# def2-TZVP/C for the RI approximation in MP2
!def2-TZVPP def2/JK RI-JK def2-TZVP/C USESYM RI-MP2
!xyzfile normalprint
%mp2
density unrelaxed # cheap unrelaxed density
natorbs true # request natural orbital (.mp2nat file)
end

* xyz 0 1
H 2.546588 -1.963946 0.000000
H -2.691521 -1.301016 0.000000
H -1.871046 -2.850177 0.000000
H 1.340164 2.755583 0.000000
H -1.161398 3.299398 0.000000
C 1.672821 -1.315704 0.000000
C -0.638005 -1.219846 0.000000
C -0.527784 0.185899 0.000000
C 0.772584 0.698710 0.000000
C -0.754442 2.295867 0.000000
N 1.920414 -0.000000 0.000000
N 0.483128 -1.958901 0.000000
N 0.598043 2.065684 0.000000
N -1.480206 1.185947 0.000000
N -1.839502 -1.841424 0.000000
*

The output contains a lot of useful information on the symmetry handling starting from
the detected point group. Note that ORCA only supports abelian point groups.
------------------
SYMMETRY REDUCTION
------------------
ORCA supports only abelian point groups.
It is now checked, if the determined point group is supported:
Point Group ( Cs ) is ... supported

(Re)building abelian point group:


Creating Character Table ... done
Making direct product table ... done
...

Thereafter, the purified geometry is printed in Cartesian and internal coordinates. Since
we requested an XYZ file the former will be saved on disk (denoted as adenine.xyz). The
output continues with the character and product tables of the irreducible
representations (IRREPs). The IRREPs are numerated as integer numbers starting with
zero e.g. A’ (0) and A’’ (1).
---------------------------
CHARACTER TABLE OF GROUP Cs
---------------------------
GAMMA O1 O2
A' : 1.0 1.0
A" : 1.0 -1.0

--------------------------------
DIRECT PRODUCT TABLE OF GROUP Cs
--------------------------------
** A' A"
70

A' A' A"


A" A" A'

Fast forwarding to the actual MP2 output, we expect the -conjugated orbitals to be of
IRREP A’’ and near the HOMO-LUMO gap. Orbitals 29-38 are indeed the desired orbitals
(Figure 22).
N[ 29]( A") = 1.98141020
N[ 30]( A") = 1.97867564
N[ 31]( A") = 1.97122770
N[ 32]( A") = 1.96035120
N[ 33]( A") = 1.94729177
N[ 34]( A") = 1.93434709
N[ 35]( A") = 0.06094152
N[ 36]( A") = 0.05502607
N[ 37]( A") = 0.04627670
N[ 38]( A") = 0.02011036

Figure 22. Adenine MP2 natural orbitals 29-38 with IRREP A''.

Reading the .mp2nat file, we proceed with the CASSCF calculation of the lowest 7 → ∗
excited states and the ground state. All these states are of the same total symmetry
(IRREP A’). In contrast to the single-reference methods, for multireference methods the
IRREP of the states can and must be explicitly specified. In ORCA, the IRREPS are
associated with a multiplicity block. In other words: Each IRREP needs a multiplicity
block. The selected active space involves 12 electrons in 10 orbitals.
!def2-TZVPP nofrozencore def2/JK RI-JK conv ri-nevpt2
!usesym moread pal8
%moinp "mp2.mp2nat" # target orbitals at the homo-lumo gap
%casscf
mult 1 #singlet states (first multiplicity block)
irrep 0 #irrep of the first multiplicity blocks
nroots 8 # number of roots for the first multiplicity block
nel 12
norb 10
end
* xyzfile 0 1 adenine.xyz

The calculation starts with low gradients (||g|| <0.3 ) and converges in 8 iterations. The
converged orbitals are denoted as cas10.gbw. The NEVPT2 excitations energies are
reported at the end of the output.
-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------
71

LOWEST ROOT (ROOT 0, MULT 1, IRREP A') = -466.649951325 Eh -12698.191 eV

STATE ROOT MULT IRREP DE/a.u. DE/eV DE/cm**-1


1: 2 1 A' 0.188913 5.141 41461.7
2: 1 1 A' 0.199764 5.436 43843.2
3: 5 1 A' 0.243292 6.620 53396.3
4: 4 1 A' 0.251591 6.846 55217.8
5: 3 1 A' 0.254247 6.918 55800.8
6: 7 1 A' 0.270466 7.360 59360.5
7: 6 1 A' 0.285344 7.765 62625.7

Next we extend the active space with the three nitrogen lone pairs to compute the two
lowest → ∗ excited states. The lone pairs must transform under A’. Lone pairs are
best identified using localized orbitals. In case of symmetry, localization must be
restricted to orbitals of the same IRREP! Using orca_loc this is a tedious approach
as the orbitals need to be rotated in place for each IRREP. However, this is automatically
done if the localization is called from the CASSCF block.
!def2-TZVPP nofrozencore def2/JK RI-JK USESYM moread
%moinp "cas10.gbw" # converged orbitals from previous calculation.
%casscf
mult 1 #singlet states (first multiplicity block)
irrep 0 #irrep A’ of the first multiplicity blocks
nroots 8 #number of roots of the first multiplicity block
nel 12
norb 10
intorbs locorbs # localize the inactive space (respecting symmetry!)
end
* xyzfile 0 1 adenine.xyz

In the last iteration of CASSCF, localized internal orbitals are produced and a summary
of the localization is printed. The orbitals are stored as cas10_loc.gbw file.
Rather strongly localized orbitals:
MO 12: 14N - 1.001056
MO 11: 13N - 0.970507
MO 10: 13N - 0.999463
MO 9: 12N - 1.000960
MO 8: 11N - 0.962805
MO 7: 11N - 0.999485
MO 6: 10N - 0.959342
MO 5: 10N - 0.999165

Visually inspecting the strongly localized orbitals we indeed find beautiful lone pairs
(Figure 23).

Figure 23. Adenine nitrogen lone pairs (localized MOs 11, 8 and 6)

Next we extend the active space with the lone pairs. For the sake of simplicity, we state-
average over the ground state, the lowest → ∗ and two → ∗ excited states.
!def2-TZVPP nofrozencore def2/JK RI-JK USESYM moread
%moinp "cas10_loc.gbw" # converged/localized orbitals
%casscf
mult 1,1 #two singlet multiplicity blocks
irrep 0,1 #irrep A’, A’’ for the two mult blocks.
nroots 8,2 #number of roots for the two mult blocks.
nel 18
72

norb 13
end
* xyzfile 0 1 adenine.xyz

The resulting SC-NEVPT2 energies with the two different active spaces are summarized
in Table 8. Using state-averaged orbitals changes the computed → ∗ by about 0.1 eV.
For more accurate results the orbitals should be optimized separately.

Table 8. Adenine vertical excitations energies in eV using state-averaged CASSCF orbitals.

Excitation Energies NEV(12,10) NEV(18,13)


2-1A’ 5.141 5.067
1A’’ - 5.46
3-1A’ 5.436 5.485
2-1A’’ - 6.089
4-1A’ 6.620 6.663
5-1A’ 6.846 6.945
6-1A’ 6.918 6.998
7-1A’ 7.360 7.303
8-1A’ 7.765 7.717

14 NEVPT2-F12 – reaching the basis set limit


The convergence behavior of the NEVPT2 correlation energy with respect to the
complete basis set (CBS) limit is similar to MP2. At a triple zeta basis set, the NEVPT2
correlation energy is typically not converged. However, explicitly correlated methods
offer an alternative approach to obtain results close to the CBS limit using a small double
zeta basis set. For RI-MP2, the methodology has been reviewed and compared against
extrapolation techniques by Liakos et al.58 Such correction have been recently developed
for the FIC-NEVPT2 approach using the RI approximation further denoted as “NEVPT2-
F12”. 59 We show how to use the NEVPT-F12 method in ORCA and compute the adiabatic
singlet-triplet gap for the Acrolein (C3H4O) molecule.

Figure 24. Acrolein CAS(6,5) selected active space.

The lowest triplet state of Acrolein is 3A”(n-π excitation). The active space should
contain 4 orbitals (with symmetry A”) and the lone pair of oxygen (with symmetry A’).
In order to get the adiabatic singlet-triplet gap, two independent CASSCF (NEVPT2)
computations have to be done for singlet and triplet geometries, respectively. Starting
point are MP2 natural orbitals as described in the previous section. Carefully choosing
the active space (Figure 24), we denote that ground and excited state guess orbitals as
“GS.mp2nat” and “ET.mp2nat”. The NEVPT2-F12 is invoked with the following input,

58 D.G. Liakos, R. Izsák, E.F. Valeev, and F. Neese, Molecular Physics 111, 2653 (2013).
59 Guo, Y.; Sivalingam, K.; Valeev, E. F.; Neese, F. to be submitted
73

where we need to specify the two F12 specific basis and auxiliary basis set. Since we are
interested into small effects, we tighten the convergence criteria.
#cc-pvdz-f12 is the F12 basis (mandatory)
#cc-pvdz-f12-cabs is the cabs auxiliary basis (mandatory)
#aug-cc-pvdz/c is chosen as RI auxiliary basis (mandatory)
!cc-pvdz-f12 aug-cc-pvdz/C cc-pvdz-f12-cabs BOHRS usesym
!moread
%moinp "GS. mp2nat"
%pal
nproc 8
end
%casscf
mult 1
irrep 0 #A’ symmetry
nel 6
norb 5
gtol 1e-5 #very tight convergence criteria
etol 1e-14
maxiter 200
orbstep nr
switchstep nr

PTMethod FIC_NEVPT2
trafostep ri
PTSettings
D4TPre 1e-14
F12 true #Enable F12 Correction
end
end
* xyz 0 1
c 1.38375499950776 0.96870923944983 0.00000000000000
o 2.59413468194069 -0.96781079895097 0.00000000000000
h 2.35864519884539 2.82377931063104 0.00000000000000
c -1.41493370061049 1.10212006319142 0.00000000000000
h -2.24310078378358 2.97813466785087 0.00000000000000
c -2.85535081360015 -0.96055402826563 0.00000000000000
h -1.99776842977842 -2.82009387269954 0.00000000000000
h -4.89779865432955 -0.85010067472844 0.00000000000000
*

We repeat the calculation for the triplet state and A’’ irreducible representation.
#cc-pvdz-f12 is the F12 basis (mandatory)
#cc-pvdz-f12-cabs is the cabs auxiliary basis (mandatory)
#aug-cc-pvdz/c is chosen as RI auxiliary basis (mandatory)
!cc-pvdz-f12 aug-cc-pvdz/C cc-pvdz-f12-cabs BOHRS usesym
!moread
%moinp "ET.mp2nat"
%pal
nproc 8
end
%casscf
mult 3
irrep 1 # A’irrep
nel 6
norb 5
gtol 1e-5 #very tight convergence criteria
etol 1e-14
trafostep ri
maxiter 200
orbstep nr
switchstep nr

PTMethod FIC_NEVPT2
PTSettings
D4TPre 1e-14
F12 true #Enable F12 Correction
end
end
* xyz 0 3
c 1.27764037014265 1.00266125484490 0.00000000000000
o 2.71859942376473 -0.97265271700710 0.00000000000000
h 2.40337112329447 2.73717394497525 0.00000000000000
c -1.35545418083404 1.07348111452171 0.00000000000000
74

h -2.12961274224224 2.96780782085672 0.00000000000000


c -2.95727209598858 -0.96045580283957 0.00000000000000
h -2.28165691696814 -2.88697749559064 0.00000000000000
h -4.97751524878229 -0.67391151257784 0.00000000000000
*

The F12 results for double zeta, triple zeta and quadruple zeta are summarized in Table
9. The absolute correlation energy deviation between the NEVPT2-F12 and the CBS limit
is less than 10 mEh at the double zeta level. The relative singlet-triplet gaps predicted by
NEVPT2-F12 are even more accurate. Clearly, with the double zeta basis set, NEVPT2-
F12 can predict very accurate singlet-triplet gap. Major deviations between NEVPT2-F12
and CBS results originate from the CASSCF calculations itself. A F12 correction for the
CASSCF might be implemented in the future.

Table 9. The absolute CASSCF and correlation energies computed by NEVPT2 and NEVPT2-F12 and the singlet-
triplet gap (in eV) with different basis sets.

State Basis set CASSCF(6,5) NEVPT2 NEVPT2-F12


cc-pvdz-f12 -190.886626 -0.592933 -0.722407
1A’ cc-pvtz-f12 -190.904917 -0.669556 -0.728607
(mEh) cc-pvqz-f12 -190.908387 -0.700183 -0.730320
CBS -190.908974 -0.730177 -

cc-pvdz-f12 -190.775675 -0.587705 -0.715330


3A” cc-pvtz-f12 -190.793803 -0.664010 -0.721774
(mEh) cc-pvqz-f12 -190.797262 -0.694047 -0.723423
CBS -190.797860 -0.723349 -

cc-pvdz-f12 3.019 3.161 3.212


Singlet-Triplet GAP cc-pvtz-f12 3.024 3.175 3.210
(eV) cc-pvqz-f12 3.024 3.191 3.212
CBS 3.024 3.209 -

15 Appendix [Cr(NH3)6]3+ - detailed walkthrough


This is a more detailed walkthrough of the example studied in Section 2, where we
explore more of the available options.

15.1 [Cr(NH3)6]3+ complex - a walk through tutorial guide


We start with a Cr(III) model complex, where in the first steps the ground state is of
interest. From elementary inorganic chemistry, we know that Cr(III) is a d3-system and
hence, the three metal d-electrons are our primary focus. In later subsections we
compute the ligand field spectrum and explore various active spaces:

· CAS(3,5) consisting of the five 3d metal orbitals.


· CAS(7,7) includes two ligand orbitals.
· CAS(7,12) extended to include the second d-shell.
75

To get familiar with the CASSCF in ORCA, we start with a small basis. The selection of the
active space orbitals, in particular the virtual 4d orbital, is easier for a small basis set. 60
This is because there are less virtual orbitals and therefore they are cleaner and without
contamination from higher lying high-angular momentum and diffuse basis functions.

For a “production level” calculation larger basis sets are required. It is possible to read
the orbitals from the converged small basis set calculation as an initial guess for a
calculation with a larger basis set. This is the recommended procedure for complicated
systems. Typically, the initial gradient is rather large but drops very quickly.

15.2 Organization of the Calculation: Preliminary Chemical Considerations

As mentioned above, the central Cr3+ ion is a d3 system. Since the [Cr(NH3)6]3+ complex is
very close to an octahedral coordination, we can expect the three metal t2g-based
orbitals to be singly occupied and hence a 4A2g ground state arises. Each NH3 provides a
single lone pair that can form a σ-bond with the central metal, but no π-bonds are
possible. Hence, we only need the two ligand orbitals of eg-symmetry when we want to
include the ligand orbitals in the active space.

For Cr, Ammonia acts as an intermediate-field ligand.61 . If we are interested in the d-d
excited states of the system, the Tanabe-Sugano diagram (Figure 25) tell us, that we
expect 4T2g and two 4T1g excited states of the same multiplicity as the ground state.62
Since T-terms are triply orbitally degenerate, 10 roots need to be calculated to capture
all spin allowed ligand field transitions. Of course, due to the presence of the hydrogens,
the actual symmetry is not strictly octahedral, but we use Oh group notation nonetheless.

Figure 25: Tanabe-Sugano diagram for d3 system in octahedral field

The initial geometry of the complex comes from a DFT geometry optimization, but it can
also be based on the crystal structure with just the hydrogens optimized.
# This is a slightly smoothed DFT optimized geometry in internal

60 Rule of thumb: Orbitals that are easy to identify also perform better as guess orbitals.
61 L.G. Vanquickenborne, B. Coussens, D. Postelmans, A. Ceulemans, and K. Pierloot, Inorg. Chem. 30, 2978
(1991).
62 Derived from the 4F and 4P terms of the free ion.
76

# coordinates. This helps keeping the orbitals clean, as the ligands will
# be placed on the coordinates. Any xyz coordinates would do too, just
# replace “int” by “xyz” and give Cartesian Coordination (in Angström)
* int 3 4
Cr 0 0 0 0 0 0
N 1 0 0 2.137 0 0
N 1 2 0 2.137 90 0
N 1 2 3 2.137 90 90
N 1 2 3 2.137 90 180
N 1 2 3 2.137 180 180
N 1 2 3 2.137 90 270
H 2 1 3 1.041 114 0
H 2 1 3 1.041 114 120
H 2 1 3 1.041 114 240
H 3 1 2 1.041 114 0
H 3 1 2 1.041 114 120
H 3 1 2 1.041 114 240
H 4 1 2 1.041 114 315
H 4 1 2 1.041 114 195
H 4 1 2 1.041 114 75
H 5 1 2 1.041 114 0
H 5 1 2 1.041 114 120
H 5 1 2 1.041 114 240
H 6 1 4 1.041 114 270
H 6 1 4 1.041 114 30
H 6 1 4 1.041 114 150
H 7 1 2 1.041 114 45
H 7 1 2 1.041 114 165
H 7 1 2 1.041 114 285
*
It is advisable to align the molecule with respect to the coordinate axis e.g. in this
example the nitrogen atoms should be along the x,y,z-axis. This merely simplifies the
identification of the orbitals later on but has no influence on the mechanics of the
calculation. An exception is the ab-initio ligand field theory (AI-LFT),63 where the
alignment of the molecule in the axis frame is important.

15.3 Initial guess CASSCF


We start with the smallest active space that is just the 3d-metal orbitals. In later
stages, we can use these converged calculations to further extend the active space. In
general, this is a good strategy for TM complexes with mostly ionic interactions
between metal and ligands, where the 3d-metal orbitals are strongly localized. For TM
complexes with delocalized 3d-metal orbitals, the overall convergence is slower if the
bonding/antibonding partners are omitted from the active space. Moreover, for such
systems, the respective ligand orbitals can be difficult to identify (see Section 6). In these
cases, it is better to directly start with an active space that comprises the 3d-metal
orbitals and their partner orbitals.

CASSCF calculations require the user to specify the number of active electron in active
orbitals. The program automatically starts with the “PModel” (~diagonalized LDA DFT
matrix) as initial guess. For many applications this is not a sufficient input as the active
orbitals are not consciously chosen. For transition metal complexes the desired 3d-metal
orbitals are often below HOMO-LUMO gap and hence do not automatically enter the
active space. Therefore, we need to carefully inspect and select the orbitals that
enter the orbital optimization. Being able to properly setup the starting orbitals is
essential for the CASSCF method. Thus, we first run the CASSCF with “NoIter” and
examine the PModel guess orbitals.
#
# Initial CASCF on [Cr(NH3)6]3+

63 Atanasov, M., Ganyushin, D., Sivalingam, K., Neese, F., Struct. and Bonding, (2012), 143:149-220
77

#
# SV = small basis set
# RI-JK = use the RI-JK approximation for the Fock matrix
# (not required – just used for more speed here)
# def2/JK = auxiliary basis for the RI approximation
# (automatically sets the auxiliary slots for
# the Fockian [AuxJK] and gradient integrals [AuxC])
# for Fock [AuxJK slot] and gradient integrals [AuxC slot])
#
# Conv = store integrals on disk (required for RIJK)
# XYZFile = leave coordinates on disk (convenient later)
# normalprint = slightly larger printing than default + includes Loewdin
# population analysis.

! SV def2/JK RI-JK Conv XYZFile noiter normalprint


%casscf
nel 3 # number of active electrons
norb 5 # number of active orbitals
mult 4 # multiplicity
nroots 1 # ground state
end

* int 3 4
Cr 0 0 0 0 0 0
N 1 0 0 2.137 0 0
N 1 2 0 2.137 90 0
N 1 2 3 2.137 90 90
N 1 2 3 2.137 90 180
N 1 2 3 2.137 180 180
N 1 2 3 2.137 90 270
H 2 1 3 1.041 114 0
H 2 1 3 1.041 114 120
H 2 1 3 1.041 114 240
H 3 1 2 1.041 114 0
H 3 1 2 1.041 114 120
H 3 1 2 1.041 114 240
H 4 1 2 1.041 114 315
H 4 1 2 1.041 114 195
H 4 1 2 1.041 114 75
H 5 1 2 1.041 114 0
H 5 1 2 1.041 114 120
H 5 1 2 1.041 114 240
H 6 1 4 1.041 114 270
H 6 1 4 1.041 114 30
H 6 1 4 1.041 114 150
H 7 1 2 1.041 114 45
H 7 1 2 1.041 114 165
H 7 1 2 1.041 114 285
*

In the following, the RI approximation is used to speed up the calculations. In Section


7.3, you will find more information on the accuracy of the RI approximation. The
keyword “xyzfile” produce a xyz coordinate file on disk that we use in later inputs
(cr_example1.xyz). The active orbitals range from 39-43. Visualization of the orbitals
helps to validate the active space. ORCA can produce cube-files that are compatible with
common visualization programs (see the manual for details). Some visualization
programs, such as Chemcraft and Avogadro, can directly read the ORCA output and plot
the orbitals if the following additional flags are set:
!Printbasis # prints the basis
%print print[p_mos] 1 end #prints MO Coefficients

A note of warning: ORCA starts the numbering of MOs and atoms from zero, while
some visualization programs start from one.64

64 Chemcraft and Avogadro start with index one.


78

In many situations, the Loewdin population analysis is sufficient for the identification. 65
Inspecting the reduced Loewdin population analysis, the 3d orbitals are readily
identified. In this example, they are perfectly in place. Note that the actual composition
printed below may differ from run to run, since some of the orbitals are degenerate
leading to arbitrary mixings.
39 40 41
-0.75958 -0.75972 -0.75991
1.00000 1.00000 1.00000
-------- -------- --------
0 Cr dxz 0.0 48.0 48.0
0 Cr dyz 0.0 48.0 48.0
0 Cr dxy 96.0 0.0 0.0

42 43
-0.37508 -0.37500
0.00000 0.00000
-------- --------
0 Cr s 0.0 0.0
0 Cr dz2 0.0 80.4
0 Cr dx2y2 80.4 0.0

In general, this is not the case and the orbitals need to be swapped manually. In
ORCA two orbitals are swapped by pair-wise Jacobi-Rotations of 90°. The following
snippet illustrates how to swap orbitals 21,22 and 23 with position 39,40 and 41.
! SV def2/JK RI-JK Conv moread
%moinp “guess.gbw” #read in the orbitals from this .gbw file

# Swap position of orbitals 21<->39, 22<->40 and 23<->41


%scf rotate {21,39,90}{22,40,90}{23,41,90} end end

%casscf
...
end
*xyzfile 3 4 “cr_example1.xyz”

Good starting orbitals have a small CASSCF orbital gradient norm at the beginning
(denoted as ||g|| in the output). They are computationally cheap and the desired active
orbitals are easy to identify. However, small initial gradients do not always translate to a
smaller number of CASSCF iterations. It is the similarity with the final orbitals that is
important. In lieu of a better quantifiable number, the gradient norm is used as a quality
measure for the initial guess. Trailing convergence in ORCA is often associated with low-
lying inactive orbitals. Although their rotations quite often dominate in the final
iterations, the resulting energy changes are fairly small.66

The default PModel guess is not necessarily the best choice. The initial gradient norm is
||g||=3.23 for this example. Here are some general remarks on other frequently used
alternatives:

· The simplest guess is “PAtom”. This gives good atomic orbitals, with an extended
Hückel like ordering, which is a good idea for transition metal calculations as the
desired metal 3d-orbitals are typically at the HOMO-LUMO gap and
essentially of metal character. For the present example, the initial gradient

65 Identification of orbitals is much easier if the molecule is in the proper axis frame!
66 We recommend using the default settings (new converger: SuperCI-PT). The SuperCI-PT is faster
compared to the older SuperCI/DIIS/NR implementations. However, the latter allow more tweaking e.g.
frozencore rotation can be disabled with the keyword “FreezeGrad” (see manual for details).
79

with ||g|| = 4.42 is larger than PModel. The PAtom guess intrinsically involves a
basis set projection. Hence, for larger basis sets, PAtom will start with an even
larger gradient.
Nevertheless, PAtom is a cheap choice to converge a CASSCF calculation
spanning the 3d-metal orbitals in a small basis. The resulting converged
CASSCF orbitals may be used to further extend the active space.
! SV def2/JK RI-JK Conv PAtom
%casscf ...
*xyzfile 3 4 “cr_example1.xyz”

· Quasi-restricted orbitals (QRO)67 from a DFT functional start with a smaller


gradient than PModel. In addition, they can be chemically interpreted on the DFT
level prior entering the CASSCF. Even for the most complicated TM-complexes
with delocalized metal orbitals or non-innocent ligands, QROs from a DFT
functional will be a fair starting point.

Note: The guess should be started with the high-spin multiplicity to avoid
convergence problems at the SCF level.
# Keyword UNO generates QROs file “.qro” on disk.
# The .qro file is a regular orbitals file (gbw).
! BP SV def2/J UNO
*xyzfile 3 4 “cr_example1.xyz”

To inspect the QRO orbitals, read the .qro file and print the orbitals in a second
run.
! BP SV def2/J moread noiter normalprint
%moinp “bp.qro” # QRO file from the previous job
*xyzfile 3 4 “cr_example1.xyz”

· Hartree-Fock orbitals (HF-QRO) of the high-spin system. 68 They are generated


using the same keyword as in the DFT calculation above.
! UHF SV def2/JK RI-JK Conv UNO
*xyzfile 3 4 “cr_example1.xyz”

Hartree-Fock orbitals are computationally more expensive compared to PAtom,


PModel and DFT orbitals. For systems with strongly localized (ionic) 3d-metal
orbitals, the similarity between HF and CASSCF orbitals reduces the number of
CASSCF iterations.
The [Cr(NH3)6]3+ complex studied here is such an example. Here, the calculations
starts with gradient norm of ||g|| = 0.01 and converges rapidly (6 iterations). For
more complicated TM complex systems, e.g. delocalized orbitals such as “ -
backbonding orbitals”, the HF orbitals are “impure” and hence difficult to identify.
The resulting guess orbitals have a large initial gradient. Similar problems occur
in organic molecules. Thus, HF-QROs are not universally recommended as
CASSCF starting orbitals.

67The construction of the QROs is described here: F. Neese, J. Am. Chem. Soc. 128, 10213 (2006).
68ROHF would be even better, but not all options are yet supported e.g. parallel run and RI will not work.
For the high-spin ground state ROHF and CASSCF are equivalent.
80

· Natural orbitals from MP2 (or higher order theory) are the most expensive guess
orbitals in this enumeration. However, the selection of “important” orbitals is
simplified with the help of the fractional occupation numbers. In particular,
bonding and antibonding partner orbitals such as − ∗ orbitals are readily
identified. Natural orbitals are an excellent choice for organic molecules. For
TM complexes, MP2 natural orbitals are not necessarily superior to the
aforementioned guess options.
# The input generates a .mp2nat file on disk.
! SV def2/JK def2-SVP/C RI-JK Conv RI-MP2
%mp2 density unrelaxed
natorbs true
end

· Inclusion of scalar relativistic effects (ZORA /DKH) strongly influences the guess
quality of PModel and PAtom and the initial gradients will be larger. If you intend
to employ ZORA or DKH in your production level calculation, then your guess
orbitals should be generated with the same corrections.

Since we start with a small basis set, a small initial gradient is not important. A few extra
iterations should be cheap. For this strategy the best choice is probably PAtom, since
it requires the least effort from the user (no inspection of orbitals) and produces
ionic orbitals, which is a very good starting point for active spaces with just the metal
orbitals. In this particular good-hearted example, no convergence problems are met.
However, as we discuss in the next section, the design of the calculation has a major
flaw!

15.4 State-Averaging as convergence aid

CASSCF optimization with active orbitals occupations close to 2.0 and 0.0 can cause
convergence problems. We discuss some of the difficulties in Section 8.4, where we
walk through a prominent example. Orbitals with this “critical” occupation are not
“correlated” and there is no benefit from including them into the active space for the
purpose of the orbital optimization. In the current example, the CAS(3,3) necessarily
converges to the same solution. Moreover, weakly/strongly occupied orbitals may be
rotated out of the active space during the orbital optimization. If the active space
composition changes by more than 50%, ORCA 4.0 will print a warning during the
iterations. Below is a fictional example snippet where the active orbitals 26 and 27 are
potentially rotated out in favor of virtual orbitals 41 and 57. The amount of mixing is
stated as OVL. Furthermore, the printing includes the overlap matrix between the active
orbitals before and after update. This matrix should be diagonal dominant with values
close to unity. Note that the numbering of active MOs starts with zero in this printing.
--- Failed to constrain active orbitals due to rotations:
Rot( 26, 41) with OVL=0.71
Rot( 27, 57) with OVL=0.64
--- SFit: squared overlap of the active orbitals QMAT=C*CT
0 1 2 3 4
0 0.050039 0.007030 -0.027847 0.000000 0.000000
1 0.007030 0.0364202 -0.000016 0.000000 -0.000000
2 -0.007847 -0.000016 0.965535 -0.000000 -0.000000
3 0.000000 0.000000 -0.000000 0.954346 -0.021121
4 0.000000 -0.000000 -0.000000 -0.031121 0.971026
...
81

The analysis is repeated in the final iteration. Here the final orbitals are compared to the
guess orbital. If the final orbitals are far away from the initial orbitals the above warning
is printed and the final orbitals should be inspected manually! If the active space truly
changed, the calculation may be restarted with re-rotated orbitals or from scratch with
different convergence settings (modified MaxRot, larger level shift or a different
converger).

To avoid convergence problems in the first place, it is advisable to do state-averaged


calculations. We have a d3 system, where the metal eg-based orbitals are unoccupied.
To populate them in the active space, we should average over a few ligand field excited
states. Thus, we have chosen to average over all 10 quartet roots (see Figure 25). The
orbitals for different d-d excited states are not expected to be very different and hence,
the only critical thing is to make sure that the eg-based orbitals have some occupation.

Here is the relevant ORCA input snippet:


%casscf nel 3 # number of active electrons
norb 5 # number of active orbitals
nroots 10 # number of roots to average over
end

Starting with the PAtom guess, the CASSCF converges within 8 iterations. The state-
average ground state energy is
ROOT 0: E= -1379.0926478359 Eh

Note that in this good-hearted example, no convergence problems are met in the state-
specific calculation.
ROOT 0: E= -1379.0936069755 Eh

State averaging came at an energetic ‘cost’ for the ground state of about 1 mEh, which is
not so bad considering that the orbitals are averaged over 10 roots.

15.5 Including ligand orbitals

In the second step, we improve the NEVPT2 reference wave function by including the
metal-ligand bonding orbitals in the active space. Having the bonding and anti-bonding
orbitals should balance the active space.

The orbitals are sorted by their energies. Hence, the desired orbitals may not be the
highest doubly occupied orbitals. In fact, they are usually not. The highest ligand orbitals
are typically non-bonding, whereas we are looking for bonding orbitals that are
stabilized. Identifying these orbitals from the previous calculation requires looking at
the orbital coefficients or better to visualize the molecular orbitals. In this example,
the reduced Loewdin analysis printed at the end of the CAS(3,5) calculation is sufficient
to identify the ligand orbitals.69 The ligand orbitals of interest are the bonding partners
of the dz2 and dx2-y2 orbitals. In our example,70 we obtain:

69 Identifying orbitals with the Loewdin analysis is bread and butter for CASSCF. A small parsing script
that filters all metal dominating orbitals might be handy.
70 Actual composition may change when repeated, since the canonical orbitals are degenerate.
82

34 35
-1.02677 -1.02677
2.00000 2.00000
-------- --------
0 Cr s 0.0 0.0
0 Cr dz2 0.0 25.2
0 Cr dx2y2 25.2 0.0
1 N s 0.7 0.2
1 N pz 0.0 0.0
1 N px 17.0 5.3
1 N py 0.0 0.0
2 N s 0.7 0.2
2 N pz 0.0 0.0
2 N px 0.0 0.0
2 N py 17.0 5.3
3 N s 0.0 0.9
3 N pz 0.0 22.3

Hence, we need to use the “rotate” feature in order to bring the correct orbitals on top
of the (formerly) doubly occupied space to include them in the active space of the next
calculation. The previously converged CAS(3,5) orbitals are denoted as “cas_5.gbw”.

Once we include the two σ-bonding orbitals of eg symmetry, the metal eg-based orbitals
obtain some population due to configuration mixing. Thus, in principle state-averaging
as a tool to help the convergence is not necessary. We nevertheless keep the 10 roots
to compute the excitation spectra.

Here is the ORCA input:


#
# Second step: include ligand orbitals
#
! SV def2/JK RI-JK conv
! moread
%moinp "cas_5.gbw" # orbitals from the CAS(3,5) calculation

# The highest doubly occupied inactive orbitals are 37 and 38.


%scf rotate {34,37,90} {35,38,90} end
end

%casscf nel 7
norb 7
mult 4
nroots 10
end

* xyzfile 3 4 cr_example1.xyz #previously generated xyzfile

This calculation converges in 9 iterations. The resulting gbw-file is denoted as


“cas_7.gbw”.

15.6 Including a second d-shell

In some cases, including a second d-shell can be useful to make the wave function more
flexible and obtain accurate results in conjunction with a subsequent second order
multireference perturbation method such as CASPT2 or NEVPT2.71 Most often it is not

71The second d-shell brings in a radial correlation effect that normally should be covered by the
dynamical correlation treatment. However, second order perturbation theory with a contracted first-
order interacting space is not flexible enough to provide this missing correlation. It is somewhat counter
83

necessary to include the entire second d-shell, but the ones that correspond to the
occupied 3d-metal orbitals.

In order to identify the second d-shell, we could localize the external orbitals from the
previous calculation. Then inspect these orbitals to identify the d-orbitals either by
visualization or using the population analysis. If necessary, reorder the orbitals such that
they are the first virtual orbitals. Note that orbitals can be localized or canonicalized
(subspace diagonalization of the Fock matrix) from the command line using the
orca_loc and orca_blockf programs – see the manual for details.

We do not recommend the aforementioned procedure, because it is tedious and quite


often results in impure virtual orbitals. The best alternative is the CASSCF option
“extorbs doubleshell”. Based on the composition of the highest active orbital, the
program automatically identifies and produces a “second shell” in the vicinity of the
active space. For an active space consisting of 3d-metal orbitals the second shell consists
of the 4d-metal orbitals. For an active space consisting of 4d-metal orbital, the second
shell consists of 5d-metal orbitals and so on.

Note that the option does not work in conjunction with symmetry (UseSym). The
following input reads the converged CAS(7,7) orbitals and produces the second-d shell
(orbitals 44-48) in the correct order.
! SV def2/JK RI-JK conv
! moread
%moinp "cas_7.gbw" # orbitals from the cas(7,7) calculation
%casscf nel 7
norb 7
mult 4
nroots 10
extorbs doubleshell # produce the double-shell above the actives.
# all other virtuals are canonicalized
end

As printed in the output, orbital nr.43 (highest active) is taken as reference and
the double-shell is produced in the MO range 44-48.
---- THE CAS-SCF GRADIENT HAS CONVERGED ----
--- FINALIZING ORBITALS ---
---- DOING ONE FINAL ITERATION FOR PRINTING ----
--- Forming Natural Orbitals
--- Canonicalize Internal Space
--- SortedExt: Largest compononent of the highest active orbital (Nr. 43) on atom 0 Cr with l=2
--- SortedExt: Double Shell Range 44 -> 48

We confirm the correctness by inspecting the Loewdin population analysis and


visualizing the orbitals (see the next subsections). The associated gbw file is denoted as
“cas_7_sorted.gbw” in the next step.
43 44 45 46 47
-0.57548 0.26599 0.26614 0.26629 0.77121
0.60166 0.00000 0.00000 0.00000 0.00000
-------- -------- -------- -------- --------
0 Cr dxz 49.4 0.0 45.2 45.2 0.0
0 Cr dyz 49.4 0.0 45.2 45.2 0.0
0 Cr dx2y2 0.0 0.0 0.0 0.0 69.3
0 Cr dxy 0.0 90.4 0.0 0.0 0.0

the philosophy of the CASSCF method (or MCSCF in general) to include dynamic correlation in the active
space. However, it is common practice and hence described here.
84

Having generated a double-shell, we will setup the calculation for the extended active
space. Since we start from an already converged CASSCF wave function, we may try the
Newton-Raphson method (keyword “switchstep nr”) to obtain convergence here. The
rate of convergence is higher with this method, but the radius of convergence is smaller.
The program can use two different convergers specified with “orbstep” and
“switchstep”. Far off from convergence “orbstep” is used. The SuperCI is good choice for
large initial gradients. ORCA changes the converger to “Switchstep” when the calculation
is close to convergence (||g|| < 0.02).72 The NR method is a safe pick for re-converging
calculations that have already been converged with a slightly different active space or
basis set.
! SV def2/JK RI-JK conv
! moread
%moinp "cas7_sorted.gbw" # cas(7,7) orbitals with prepared virtual space.

%casscf nel 7
norb 12 #3d + ligands + 4d orbitals
mult 4
nroots 10
cistep accci # faster, more memory hungry algorithm for the CI step
switchstep nr
end

* xyzfile 3 4 cr_example1.xyz

In many cases, switching to the computationally more demanding NR solver does not
result in net time savings. In this example, the “switchstep NR” and the default converger
perform equally well (4-6 iterations).
For larger active spaces or many roots, the timings can be considerably improved using
the “CIStep ACCCI” for the CI calculation. The method is absolutely equivalent to the
default CI solver, but uses are more memory demanding algorithm. The final set of
orbitals is denoted as “cas_12.gbw” in the next section.

15.7 Basis set projection (increasing the basis set)

Small basis sets have their advantages in the selection of the active orbitals. In this step,
we take the target basis set (in this case, a relatively large and flexible def2-TZVPP
basis) and re-converge the calculation. Note that the basis-set handling/naming
convention has changed in ORCA 4.0 (see the manual for details). The auxiliary basis
“def2/JK” is equally suited for the larger def2-TZVPP basis.

In ORCA, the basis set projection is enforced by simply reading the orbitals.
#
# Scale up the basis set
#
# def2/JK is a big fitting basis, good for any purpose
#
! def2-TZVPP def2/JK RIJK conv
! moread
%moinp "cas_12.gbw" # converged CAS(7,12) in SV basis

%casscf nel 7
norb 12

72 controlled by the keyword “switchconv”


85

mult 4
nroots 10
cistep accci
end

* xyzfile 3 4 cr_example1.xyz

The inactive and active orbitals are independently least-square fitted to minimize the
overlap with input orbitals. Nevertheless, the initial gradient is rather large. For the
convergence, it is a critical step as it involves the largest set of (orbital) rotation
parameters.

Aside from the energy and the gradient, ORCA prints a lot of information during the
iterations. This information might be handy when running into convergence problems.
The occupation numbers printed during the iteration are not necessarily natural orbital
occupations.73 By default the respects the initial representation of the active orbitals e.g.
localized orbitals stay localized during the iterations.
E(CAS)= -1379.331962665 Eh DE= 0.000000000
--- Energy gap subspaces: Ext-Act = -1.101 Act-Int = 0.054
--- current l-shift: Up(Ext-Act) = 2.70 Dn(Act-Int) = 1.65
N(occ)= 1.97861 1.97858 0.59957 0.59956 0.59956 0.61545 0.61544 0.00286 0.00286 0.00286
0.00233 0.00233
||g|| = 7.226001358 Max(G)= 2.391066084 Rot=499,16

Notice that ORCA prints the energy separation between external, active and inactive
spaces - ideally this number is 0.2 or larger. Convergers may fail or may not preserve the
active space when the energy separation becomes negative. Therefore, ORCA by default
employs a dynamical Level-shifting to disentangle the three subspaces. Level shift is the
main lever to influence convergence of the SuperCI, DIIS and NR approaches.74 In our
own experience, with an appropriate level-shift the combination of “orbstep SuperCI”
and “switchstep DIIS” scheme is very robust.

The default converger (SuperCI_PT) is more aggressive and operates without level
shifts. Here, “MaxRot” is the parameter that adjusts the stepsize.75 It is an empirical
parameter, and might need some fine-tuning depending on the system. For example,
when facing trailing convergence, unrestricting MaxRot is a good idea (MaxRot >1).
However, if the guess active orbital are “impure”, reducing MaxRot helps to preserve the
active space.

In this example, the gradient norm starts with ||g||=7 and it takes 8 iterations to meet
the convergence criteria (||g|| <10-3 or DE<10-8).

The orbitals and the occupation numbers from the converged calculation are shown
below. They are ordered by increasing occupation number. You see an ideal shape and
ordering of the orbitals, which makes the interpretation of the results very convenient. It
is also a quality control for your calculation to ensure that you have arrived at the
desired enlarged active space. Note that not all of the orbitals are perfectly aligned to the
coordinating ligands. This is perfectly normal as some of the orbitals are degenerate and
hence arbitrarily mixed.

73 Can be enforced with the keyword “actconstraints 2” in the CASSCF block, but is not recommend.
74 Either a constant level shift (“ShiftUP”/ “ShiftDN”) or a larger dynamical shift (“MinShift”)
75 keyword Ma(default = 0.2).
86

Figure 26: These two orbitals are the antibonding eg-counterparts in the second d-shell. Notice how large
these orbitals are. If we would plot a radial cut, you would observe that they have a node, whereas the
primary d-orbitals do not have it. (isosurface value 0.05)

Figure 27: These three orbitals are the second d-shell counterparts of the nonbonding t 2g based metal
orbitals.

Figure 28: These three orbitals are the nonbonding metal d-orbitals of t 2g origin.

Figure 29: These are the antibonding eg based orbitals in the primary metal d-based set.
87

Figure 30: These two orbitals are the essentially doubly occupied bonding counterpart of the metal eg-
orbitals

The resulting gbw-file with the converged orbitals is denoted as “cas_12.gbw”.

15.8 Calculating the ligand field spectrum

So far we have considered only quartet terms. Next, we add the spin-doublet roots and
dynamical correlation by adding a SC-NEVPT2 correction on top of the CAS.76 According
to the Tanabe-Sugano diagram in Figure 25, the next double states are derived from the
free ion 2G term, which splits into the triply degenerate 2T2, 2T1, the doubly degenerate
2E terms and an 2A term. Thus, we need a total of nine doublet roots.
1

Here is the ORCA input for the most complete calculation involving the doublets,
quartets and the large active space:
#
# calculate the ligand field spectrum
#
! def2-TZVPP def2/JK RI-JK conv
! moread
%moinp "cas_12.gbw" #previously converged CASSCF calculation

%casscf nel 7
norb 12
mult 4,2 # quartet and doublet multiplicities
nroots 10,9 # 10 quartets, 9 doublets
# you can adjust the weight of each manually
# as described in the manual. Default is equal
# weights for multiplicity blocks.
PTMethod SC_NEVPT2 # invoke the SC-NEVPT2 correction
end

* xyzfile 3 4 cr_example1.xyz

In ORCA 4.0, the default state-averaging sets equal weights for multiplicity blocks.
The actual weights are also printing in the output before the first CASSCF iteration, when
the CI is setup. Let us look at some of the results of this calculation. After convergence,
you find the definition of the wave function for each state:
---------------------------------------------
CAS-SCF STATES FOR BLOCK 1 MULT= 4 NROOTS=10
---------------------------------------------

ROOT 0: E= -1379.8845593013 Eh
0.96625 [ 0]: 221110000000
0.00458 [ 2552]: 111111100000

76 NEVPT2 does not yet support the faster ACCCI as CIStep.


88

0.00440 [ 165]: 211111000000


0.00439 [ 1828]: 121110100000
0.00271 [ 809]: 201112000000
0.00271 [ 6660]: 021110200000
ROOT 1: E= -1379.7943086004 Eh 2.456 eV
0.73142 [ 1]: 221101000000
0.23945 [ 9]: 221010100000
0.00410 [ 2]: 221100100000
0.00271 [ 1835]: 121101100000
...

The program then lists the main contributing configurations that are active space
occupation patterns. For example, the ground state has a weight of 0.96625, which
means that the lowest root is dominated to 96.6% by a single configuration (this number
is the sum of the squares of the CI coefficients for all configuration state functions that
belong to this configuration, e.g. the linearly independent spin couplings). This
configuration has the active space occupation pattern 221110000000 which means the
first active orbital is doubly occupied, the second doubly occupied as well, the next three
orbitals are singly occupied and the remaining orbitals are empty. If you look at your
orbitals (see above), you see that the first two are the ligand based bonding orbitals, the
next three the metal t2g based orbitals, followed by the two metal eg based orbitals and
the remaining ones are the second d-shell.77 ORCA by default uses natural orbitals for
the active space. The metal eg based orbitals have a slightly higher occupancy due the
presence of the ligand orbitals.

Note that ORCA employs configuration state functions. Occasionally one interested in
the CI Coefficients or the representation in terms of spin determinants. This is
possible with the keyword “PrintWF” and discussed in Section 8.5 in more detail.

The program then prints the CASSCF transition energies:


-----------------------------
SA-CASSCF TRANSITION ENERGIES
------------------------------

LOWEST ROOT (ROOT 0 ,MULT 4) = -1379.884565085 Eh -37548.568 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 0 2 0.086665 2.358 19020.7
2: 1 2 0.086672 2.358 19022.4
3: 2 2 0.090109 2.452 19776.6
4: 3 2 0.090113 2.452 19777.5
5: 4 2 0.090115 2.452 19777.9
6: 1 4 0.090270 2.456 19812.0
7: 2 4 0.090279 2.457 19813.9
8: 3 4 0.090295 2.457 19817.5
9: 5 2 0.125745 3.422 27597.8
10: 6 2 0.125752 3.422 27599.5
11: 7 2 0.125760 3.422 27601.1
12: 4 4 0.128986 3.510 28309.2
13: 5 4 0.128987 3.510 28309.4
14: 6 4 0.129017 3.511 28316.0
15: 8 2 0.160764 4.375 35283.6
16: 7 4 0.202961 5.523 44544.8
17: 8 4 0.202968 5.523 44546.3
18: 9 4 0.202991 5.524 44551.4

For convenience, three different units are printed (au, eV and cm-1). What follows is a
population analysis and orbital printing before we reach the NEVPT2 section:

77 The number in square brackets is the number of the configuration in the configuration list and is
irrelevant.
89

----------------------------------------------------------------
< NEVPT2 >
----------------------------------------------------------------

SYSTEM-SPECIFIC SETTINGS:
Number of active electrons ... 7
Number of active orbitals ... 12
Total number of electrons ... 81
Total number of orbitals ... 502
Total number aux. functions ... 1050

There is a fair amount of output generated in the course of the calculation that, most of
the time, is of limited interested to the user. However, eventually we reach the section:
===============================================================
NEVPT2 Results
===============================================================

For the really curious user the program then prints the contributions of each excitation
class to the final NEVPT2 correction. Finally, we obtain:
-----------------------------
NEVPT2 TRANSITION ENERGIES
------------------------------
LOWEST ROOT (ROOT 0, MULT 4) = -1381.655526090 Eh -37596.758 eV

STATE ROOT MULT DE/a.u. DE/eV DE/cm**-1


1: 1 2 0.075667 2.059 16606.9 #2Eg
2: 0 2 0.075713 2.060 16617.1
3: 2 2 0.078498 2.136 17228.3 #2T1g
4: 3 2 0.078502 2.136 17229.2
5: 4 2 0.078569 2.138 17244.0
6: 3 4 0.096120 2.616 21095.8 #4T2g
7: 1 4 0.096178 2.617 21108.6
8: 2 4 0.096188 2.617 21110.8
9: 7 2 0.113633 3.092 24939.5 #2T2g
10: 5 2 0.113635 3.092 24940.1
11: 6 2 0.113641 3.092 24941.3
12: 4 4 0.128484 3.496 28198.9 #4T1g
13: 5 4 0.128496 3.497 28201.7
14: 6 4 0.128515 3.497 28205.8
15: 8 2 0.157158 4.276 34492.1 #2A1g
16: 7 4 0.207238 5.639 45483.5 #4T1g
17: 8 4 0.207339 5.642 45505.7
18: 9 4 0.207367 5.643 45511.7

These are the final transition energies. We see that the lowest state is a quartet (as
expected) and the next higher states are doublets starting at around 17000 cm-1. Hence,
we have an energetically well-isolated ground state. The states can be readily assigned
with the aid of their degeneracy and the Tanabe-Sugano diagram for d3 as indicated
above.

For convenience, the results of the aforementioned calculations are summarized Table
10. It is evident that the changes from CASSCF to NEVPT2 are not enormous and
amount approximately 0.2 eV. This indicates that the CASSCF description of the
spectrum is already pretty good and the NEVPT2 results are reliable. Another
interpretation can be that static electron correlation is dominating in this Cr complex,
hence the recovered dynamic electron correlation doesn’t change the result much.
Indeed, as far as comparison to experiment is possible, the results are within 0.3 eV. This
is a good result given that the relativistic effect was neglected, environment effects not
included and no attempt has been made to reach the basis set limit.

Table 10. Few energy ligand field spectra using the default weighting (equal weights for multiplicity blocks).
The CAS(7,12) consists of the 3d orbitals, 2 ligand orbitals and the second d-shell.
90

State Exp.a CASSCF NEVPT2


(7,12) (7,12)
2E 15300 19020 16606
g
2T 15300 19776 17228
1g
2T - 27597 24939
2g
4T 21550 19812 21095
2g
4T 28500 28309 28198
1g
2A - 35283 34492
1g
4T - 44544 45483
1g
a Jorgensen, C., K., Absorption Spectra and Chemical Bonding in Complexes, Pergamon Press, Oxford, 1962,

p291 and references therein.

Let us investigate the influence of state-averaging and the extension of the active space
on the ligand field spectrum computed with NEVPT2. For comparison, we provide the
following results in Table 11:
· Results with the minimal SA-CAS(3,5) that is just 3d-metal orbitals. The orbitals
are optimized for the quartet states
· SA-CAS(3,5) with the default weighting for the orbital optimization.
· Results with two ligand orbitals included in the active space: SA-CAS(7,7)
· Results with the complete second d-shell included: SA-CAS(7,12).

Averaging over just the quartets or the quartets and the doublets has a minor effect
(<0.05 eV), which is expected as the optimal orbitals for all ligand field states are
expected to be fairly similar. In this example, inclusion of the ligand orbitals barely
changes the spectra. The effect of ligand orbitals is more important for covalent ligands.
The effect of the second d-shell is recognizable but not overwhelmingly large. The
second d-shell is more pronounced for complexes, where the 3d orbitals are strongly
occupied.

Table 11: NEVPT2 Vertical Transition energies of [Cr(NH3)6]2+ in cm-1. The calculation involves 10 quartet
and 9 doublet states. CAS(7,7) includes ligand orbitals. CAS(7,12) includes ligand orbitals and the second d-
shell.

State NEPT2 NEVPT2 NEVPT2 NEVPT2 Exp.c


CAS(3,5)a CAS(3,5)b CAS(7,7)b CAS(7,12)b
2E 17117 17056 17194 16606 15300
g
2T 18001 17926 17876 17228 15300
1g
2T 25966 25878 25761 24939 -
2g
4T 20360 20596 21033 21095 21550
2g
4T 28847 29019 28440 28198 28500
1g
2A 34370 34534 34970 34492 -
1g
4T 43991 44439 45238 45483 -
1g
a Orbitals optimized for the quartet states.
b Orbitals optimized with equal weights for quartets and doublets.
c Jorgensen, C., K., Absorption Spectra and Chemical Bonding in Complexes, Pergamon Press, Oxford, 1962,
p291 and references therein.

You might also like