Autumn School on Correlated

Effective Hamiltonians: Theory and Application

Frank Neese, Vijay Chilkuri, Lucas Lang

MPI für Kohlenforschung Kaiser-Wilhelm Platz 1 Mülheim an der Ruhr What is and why care about effective Hamiltonians? Hamiltonians and Eigensystems

★ Let us assume that we have a Hamiltonian that works on a set of variables x1 .. xN. H x ,..., x ( 1 N ) ★ Then its eigenfunctions (time-independent) are also functions of x1 ... xN.

H x ,..., x Ψ x ,..., x = E Ψ x ,..., x ( 1 N ) I ( 1 N ) I I ( 1 N )

★ The eigenvalues of the Hamiltonian form a „spectrum“ of eigenstates that is characteristic for the Hamiltonian

E7 There may be multiple E6 systematic or accidental E4 E5 degeneracies among E3 the eigenvalues

E0 E1 E2 Effective Hamiltonians

★ An „effective Hamiltonian“ is a Hamiltonian that acts in a reduced space and only describes a part of the eigenvalue spectrum of the true (more complete) Hamiltonian

E7

E6 Part that we want to E4 E5 describe E3

E0 E1 E2

TRUE H x ,..., x Ψ x ,..., x = E Ψ x ,..., x (I=0...infinity) ( 1 N ) I ( 1 N ) I I ( 1 N ) (nearly) identical eigenvalues wanted!

EFFECTIVE H Φ = E Φ eff I I I (I=0...3) Examples of Effective Hamiltonians treated here

1. The Self-Consistent Size-Consistent CI Method to treat correlation ‣ Pure ab initio method. ‣ Work in the basis of singly- and doubly-excited determinants ‣ Describes the effect of higher excitations

2. The -Hamiltonian in EPR and NMR Spectroscopy. ‣ Leads to empirical parameters ‣ Works on fictitous effective electron (S) and nuclear (I) spins ‣ Describe the (2S+1)(2I+1) ,magnetic sublevels‘ of the electronic ground state

3. The Heisenberg-Dirac-van Vleck Hamiltonian of Molecular ‣ Empirical, parameterized model ‣ Works on fictitious electron spin variables of ,magnetic subsystems’; ‣ Describes a few low-lying multiplets

4. The Ligand Field Hamiltonian of coordination chemistry ‣ Empirical, parameterized model ‣ Works on a fictitious set of n-electrons in the d- or f-orbitals of a d- or f-element ‣ Describes d-d excited states in transition metal complexes Other Examples of Effective Hamiltonians

1. The Hubbard Hamiltonian of Molecular Magnetism ‣ Works on fictitious single sites ‣ Refinement of the HDvV Hamiltonian with ,on-site‘ electron repulsion

2. The ,Double Exchange‘ Hamiltonian of mixed valence systems and the ,Electron transfer‘ Hamiltonian of electron transfer theory ‣ Describes only a two site system with localized electrons

3. The Hückel Hamiltonian for aromatic systems ‣ Describes π-Electron excited states

4. The Quasi-Relativistic ZORA Hamiltonian ‣ Describes the ,large component‘ of the spinor

5. … (hundreds of uses throughout chemistry and physics) The Value of Effective Hamiltonians

✓ EH’s are MUCH simpler than the ‚parent‘ Hamiltonians ‣ Treat their eigensystems analytically or with little effort numerically ‣ Help to Identify the minimum number of physically sensible empirical parameters to effectively describe the physical situation at hand.

✓ EH’s have a great imaginative power: ‣ They create pictures in which we can think ‣ They provide a language in which we can talk ‣ They provide insights into classes of substances rather than numbers for individual systems

➡ GOOD effective Hamiltonians have parameters that have an unambiguous definition in terms of first principle physics

➡ LESS GOOD effective Hamiltonians have parameters with a cloudy of ill defined connection to first principle physics

Following this logic, the Spin Hamiltonian is a GOOD effective Hamiltonian while the Hückel Hamiltonian is a less good effective Hamiltonian. What do we mean by the „Complete Hamiltonian“?

(with a short excursion into ) What is the „Complete Hamiltonian“ ?

‣ One could always take the many particle four component relativistic Hamiltonian with inclusion of external electric and magnetic fields as complete Hamiltonian as it describes, to the best of our knowledge, all chemical phenomena

‣ For EPR (in particular!) and NMR (to some extent) theory this is essentially the case. Hence, the EPR physics is a very complete one!

‣ For many other effective Hamiltonians it is enough to regard the Born- Oppenheimer Hamiltonian as ,the complete Hamiltonian‘; for example in the H-D- vV Hamiltonian.

‣ Sometimes even an effective Hamiltonian at one level may serve as the basis for an effective Hamiltonian at another level (e.g. the full spin Hamiltonian in relation to the description of only a nuclear spin manifold) The Molecular Hamiltonian

e- T e- Vee e

VeN Te VeN

Hˆ( , )Ψ( , ) = EΨ( , ) E = Total VeN VeN x R x R x R N+ VNN N+

Hˆ(r, R) = T +T +V +V +V e N eN ee NN 2 2 e0 ZA 2 e Z Z V = − T = p / 2m 0 A B eN ∑ e ∑ i el VNN = ∑ 4πε0 A,i r − R i 4πε A

Note: i ,j sum over electrons, A,B sum over nuclei, ZA,MA- Charge and Mass of Nucleus ,A‘ are Many Particle Systems

Schrödinger equation ˆ HΨ(r1 σ1 ,...,rNσN ) = EΨ(r1 σ1 ,...,rNσN )

Hamiltonian ˆ ˆ ˆ ˆ H = HBO + Hrel + HFields

SM No more good Ψ Quantum numbers

Many particle Spectrum E ,E ,.....,E That is the only thing that is seen in 1 2 ∞ experiments that probe energy

Hartree-Fock/DFT ˆ ˆ H(r1 σ1 ,...,rNσN ) → F(r1 σ1 ) ˆ Fψi (r1 σ1 ) = εi ψi (r1 σ1 )

Orbitals, NOT States! Complete different from many particle One particle Spectrum ε1 ,ε2,.....,ε∞ spectrum; NOT observable; different for each approximated state HS-DFT S=9/2 S=7/2 S=5/2 S=3/2 S=1/2 BS-DFT

➡The particle/hole spectrum is very far from the real spectrum! Extreme Example: L-Edge X-Ray Absorption

L-edge excitations lead to final states with a (n-1)p core hole, e.g.:

nd X-Ray

(n-1)p

Electronic State:

6 x 25 6Γ,4Γ, 2Γ = 550 CSFs (1512 STATES) only 15 particle/hole pairs

The p/h space does only spans a small part of the final state manifold! Its eigenspectrum cannot possibly coincide with the true (non-relativistic) eigenspectrum Numerical Data: The Mn2+ Ion (6S, high-spin d5)

Many particle spectrum p/h spectrum (CI) (BP86 TD-DFT +30 eV)

This looks and actually is unrealistic

Focusing only on the sextet states the p/h pattern is qualitatively correct but all multiplets arising from spin-flips are missing.

6P

6 Transition Energy (eV) Energy Transition D 6F However, once SOC is 6P turned on the grave 4P 6D deficiencies of the 6 F p/h spectrum will show up Effective Hamiltonians through Partitioning Theory Expansion of the Wavefunction

✓ Assume that we have defined our „Complete Hamiltonian“. Assume that we can (or

should) divide it into a part H0 and a part H1

H Ψ = (H (0) + H (1))Ψ = E Ψ I I I I

✓ The solutions to H (0)Ψ(0) = E (0)Ψ(0) I I I

✓ are assumed to be known (BOLD assumption!)

✓ Then we can always expand the eigenfunctions of the full Hamiltonian in terms of the eigenfunctions of the 0th order Hamiltonian:

Ψ = C Ψ(0) I ∑ JI J J

✓ Hence, the Schrödinger equation turns into a matrix eigenvalue problem

HC = EC The Partitioning Approach

✓ Critical step: divide the 0th order states into the ,a‘ set that (=model space; the functions that dominate the final states of interest - very small!)

✓ The b-space or ,outer space‘. The outer space can be very large!

✓ Partitioned eigenvalue problem:

⎛ AA AB ⎞⎛ A ⎞ ⎛ A ⎞ ⎜ H H ⎟⎜ C ⎟ ⎜ C ⎟ ⎜ ⎟⎜ ⎟ = E ⎜ ⎟ ⎜ BA BB ⎟⎜ B ⎟ ⎜ B ⎟ ⎜ H H ⎟⎜ C ⎟ ⎜ C ⎟ ⎝ ⎠⎝ ⎠ ⎝ ⎠ ✓ The equation for the ,b‘ space coefficients can be formally solved:

CB = −(HBB −1E)−1 HBACA

✓ Hence:

HAACA − HAB(HBB −1E)−1 HBACA = ECA

Expansion of the Partitioned Eigenvalue Problem

✓ Hence: Heff (E)CA = ECA

✓ With the effective Hamiltonian:

Heff (E) = HAA − HAB(HBB −1E)−1 HBA dimension=dim(A)xdim(A)

✓ Exact equation!

✓ However, since the desired energy E is contained in the effective Hamiltonian, the equation is nonlinear and difficult to solve.

✓ We will pursue a simple approach here that exposes the nature of the reasoning. First let us look at the Hamiltonian in b-space:

BB (0) (0) (0) (0) (1) (0) (H )IJ = ΨI | H | ΨJ + ΨI | H | ΨJ = δ E (0) + Ψ(0) | H (1) | Ψ(0) Coupling among the b-space IJ I I J eigenfunctions can be neglected if (0) ≈ δ E H1 is much smaller than H0 IJ I Simplification of the Effective Hamiltonian

✓ Realize that we seek solutions in the vicinity of the eigenvalues of HAA - possible if the coupling to the b-space is not too large.

✓ Dropping this restriction leads to the reasoning of Malrieu‘s intermediate Hamiltonians that contain a ,buffer space‘ to ,protect‘ the model space against strong perturbers.

✓ With that assumption, we can replace:

(1E) ≈ δ E IJ IJ a again neglecting the ,small‘ coupling of the ,a‘ 1 (0) E = E states via H1 (but we could have taken a ∑ I dim(A) I ∈'a ' eigenvalues of H0+H1 in ,a‘ space equally well

✓ Then we are done Heff = HAA − HAB(EBB −1E)−1 HBA

Matrix Elements of the Effective Hamiltonian

eff (0) (0) (1) (0) (H )IJ = δIJ EI + ΨI | H | Ψj Ψ(0) | H (1) | Ψ(0) Ψ(0) | H (1) | Ψ(0) − I K K J ∑ (0) E − E K ∈'b ' K a

✓ This looks like second order perturbation theory but is more general since the

coupling of the ,a‘ space functions via the perturbing operator H1 is taken into account.

✓ We could have arrived at this result as well by a formal series expansion of the inverse matrix that would then also define higher order corrections to the effective Hamiltonian but for most intents and purposes the second order Heff is the desired one. Summary

✓ In order to apply the effective Hamiltonin theory in the proposed form the following conditions have to be met:

1. There must be a sensible division of the ,Complete Hamiltonian‘ into H0

and H1.

2. We must know the complete set of eigenfunctions of H0 3. There must be a large enough energy gap between the model space and

the outer space (Hence, the matrix elements of H1 should not be so large as to induce a crossing or near crossing of the b-space eigenfunctions with the ,a‘ space eigenfunctions).

✓ All three asusmptions may or may not be critical. In particular (2)+(3) are sometimes hard to meet and then one has to look into an alternative approach (→linear response theory) Example 1: The Heisenberg Hamiltonian

„Parameterizing a few low-lying electronic states“ What is Exchange ?

The interaction of two paramagnetic ions (or more generally fragments) leads to a „ladder“ of total spin states which are described phenomenologically by the Heisenberg-Dirac-VanVleck Hamiltonian

St=4

Ion A Ion B 8J

St=3

N N 6J SA=5/2 SB=3/2 St=2 S S 4J Partially Filled Net Net Partially Filled „Magnetic“ St=1 Magnetic Magnetic d-Shell Interaction d-Shell Moment Moment 2J

St=0

What is the origin of this „magnetic“ interaction and how do we calculate it?

With no other magnetic interactions, the energy of a given spin-state is simply: Effective Hamiltonian Treatment of the Heisenberg Model

✓ In order to derive the Heisenberg Hamiltonian in the simplest case (the Anderson model). we make the following specification of the general second-order effective Hamiltonian

1. H0 is the Epstein-Nesbet Hamiltonian (diagonal of the CI matrix) and H1 =

H - H0. Thus, the complete Hamiltonian is the Born-Oppenheimer Hamiltonian. 2. This means, we do know the eigenfunctions of the 0th order Hamiltonian exactly (Slater determinants). 3. Our model space for the most elementary case of two interacting S=1/2

systems consists of two ,neutral‘ determinants |core(aαbβ)> and |

core(aβbα)>. 4. We assume that we know the quasi-localized orbitals ,a‘ and ,b‘. 5. The outer-space consists of all other Slater determinants including the

ionic ones |core(aαaβ)> and |core(bαbβ)> and we restrict attention to those Evaluation of the Effective Hamiltonian

★ Assuming two (semi) localized orbitals ,a‘ and ,b’, then the model space is:

(core)a b (core)a b α β β α ★ The ,+‘ and ,-‘ combination of these determinants are the M=0 components of the lowest singlet and the lowest triplet respectively. ★ The diagonal elements of the effective Hamiltonian are equal for both model functions and hence may be put to 0.

★ The off-diagonal first order term is: a b | H | a b = (a b | a b ) = K α β 1 β α α α β β ab ★ And the off-diagonal second-order term is:

a b | H | a a a a | H | a b a b | H | b b b b | H | a b − α β 1 α β α β 1 β α − α β 1 α β α β 1 β α a a | H | a a − a b | H | a b b b | H | b b − a b | H | a b α β 0 α β α β 0 α β α β 0 α β α β 0 α β Extraction of the Exchange Coupling Constant

✓ Since: aαbβ | H1 |aαaβ = aαbβ | H1 |bαbβ = Fab

aαaβ | H 0 |aαaβ − aαbβ | H 0 |aαbβ = (aαaα |aβaβ )−(aαaα |bβbβ ) = Jaa −Jab ≡U

✓ We obtain the effective Hamiltonian: ⎛ 2 ⎞ ⎜ F ⎟ ⎜ ab ⎟ eff ⎜ 0 K − 2 ⎟ H = ⎜ ab U ⎟ ⎜ ⎟ ⎜ ⎟ ⎝ cc 0 ⎠ ✓ And the splitting: 2 Fab E(S = 0)− E(S = 1) = 2Kab − 4 U ✓ And from the Spin-Hamiltonian HHDvV = −2JSASB = −J(S 2 −S 2 −S 2 ) A B E(S = 0)− E(S = 1) = 2J ✓ Hence 2 Fab J = Kab − 2 U 3+ A Model Calculation: [Cu2(µ-F)(H2O)6]

The Hartree-Fock SOMOs of the triplet state („active“ orbitals)

~0.7 eV

The pseudo-localized „magnetic orbitals“

Notes: • ‚a‘ and ‚b‘ have tails on the bridge (and on the other side) • ‚a‘ and ‚b‘ are orthogonal and normalized • ‚a‘ and ‚b‘ do not have a definite energy • THE orbitals of a compound are not well defined! (ROHF, MC-SCF, DFT, Singlet or Triplet Optimized, ...) Values of Model Parameters:

„Direct“ (Potential) exchange term:

Exactly calculated „kinetic“ exchange term:

Is that accurate? Look at the singlet wavefunction: Far off

BUT: • The ionic parts are too high in energy and mix too little with the neutral configuration (electronic relaxation) • Need to include dynamic correlation into the calculation

Recommended Literature: Calzado, C. J.; Cabrero, J.; Malrieu, J. P.; Caballol, R. J. Chem. Phys. 2002, 116, 2728 Calzado, C. J.; Cabrero, J.; Malrieu, J. P.; Caballol, R. J. Chem. Phys 2002, 116, 3985 Fink, K.; Fink, R.; Staemmler, V. Inorg. Chem. 1994, 33, 6219 Ceulemans, A.; et al., L. Chem. Rev. 2000, 100, 787 Refined Ab Initio Calculation

Include relaxation and LMCT/MLCT states via Difference Dedicated CI: (~105 Configurations)

Look at the singlet wavefunction

Reduced Increased! NEW+IMPORTANT The Anderson model is not really realistic and should not be taken literally even though its CI ideas are reasonable. ‣ Relaxation of ionic configurations are important („dressing“ by dynamic correlation) ‣ LMCT states are important

Treatment of LMCT States in Model Calculations: VBCI Model: Tuczek, F.; Solomon, E. I. Coord. Chem. Rev. 2001, 219, 1075 Comments

✓ Kab is always positive (ferromagnetic). „Potential Exchange“

✓ -F2/U is always negative since Jaa > Jab (antiferromagnetic). „Kinetic Exchange“

✓ This effective Hamiltonian is too simple and upon ab initio evaluation of the integrals one recovers only a fraction of J.

✓ The reason is that the ,bare‘ U is much too large since the ionic configurations relax a lot in the dynamic correlation field.

✓ The dynamic correlation contributions can - again - be calculated through an effective Hamiltonian.

de Loth, P.; Cassoux, P.; Daudey, J. P.; Malrieu, J. P. J. Am. Chem. Soc. 1981, 103, 4007; Calzado, C. J.; Cabrero, J.; Malrieu, J. P.; Caballol, R. J. Chem. Phys. 2002, 116, 3985; Calzado, C. J.; Cabrero, J.; Malrieu, J. P.; Caballol, R. J. Chem. Phys. 2002, 116, 2728. Miralles, J.; Caballol, R.; Malrieu, J. P. Chem. Phys. 1991, 153, 25; Miralles, J.; Daudey, J. P.; Caballol, R. Chem. Phys. Lett. 1992, 198, 555. A Comment on Broken Symmetry

Simplest possible case

1

1/2

1 2 a = 1 ψ + ψ b = 1 ψ − ψ 2 ( g u ) 2 ( g u ) Ψ00 = 1 ab − ab 2 ( ) ✓ tails on the bridge, orthogonal

Terrible mistake: In the real world Broken Symmetry DFT: everywhere zero for a singlet state!!! BS Ψ = ηaηb Positive spin Negative spin density ✓ Energy gain through delocalization,density non-orthogonal

My View: FN (2003) J. Phys. Chem. Solids, 65, 781; FN Coord. Chem. Rev., 2009, 253,526 Example 2: The Spin-Hamiltonian

„Incorporating additional small effects (here relativistic and external field effects)“ Magnetic Interactions

IB1QB1IB1 IA1QA1IA1

β Bg I βNBgNB1IA1 IA1 N NA1 A1 B IB1 β e Bg S A S A B S A A Bg B β e B1 A1 I

I B1 I A1 A B1 B

S J NMR A2 I

S S I

A B B2 NMR SAJSB J

A1 S I B A I A2 SADASA SBDBSB B2 I A2 B2 A S A J ~ 0 -103 cm-1 D ~ 10-1-101 cm-1 I -1 1 -1 I A2 βeg ~ 10 -10 cm B2 A ~ 0 -10-2 cm-1 -2 -1 IA2QA2IA2 βNgN ~ 0 -10 cm IB2QB2IB2 Q ~ 0 -10-3 cm-1 -8 -1 JNMR ~ 0 -10 cm The Spin Hamiltonian: Summary

! ! 2 1 E 2 2 Hˆ = SDS (= D(S − S(S + 1)) + (S −S )) Zero-Field Splitting Spin ! ! z 3 D x y + βBgS ! ! Zeeman Term (g-Tensor) + SA(A)I (A) ∑ Hyperfine Interaction A ! ! (A) (A) (A) + ∑I Q I Quadrupole Interaction A ! ! − B (A)I (A) Nuclear Zeeman βN ∑ gN A ! ! + ∑ I (A)J(A,B)I (B) Spin-Spin Coupling A

! S Fictitous Electron Spin β Bohr‘s Magneton ! I (A) Nuclear Spin β Nuclear Magneton N Theoretical Magnetic Spectroscopy

Molecular Hamilton-Operator

Molecular- Direct Calculation structure

Theory

Spectra Spin Hamilton-Operator Fit

Simulation

Neese, F. Quantum Chemistry and EPR Parameters eMagRes 2017, 6, 1. (and many other reviews since 2001) Reviews: (1) FN Curr. Op. Chem. Biol., 2003, 125 Effective Hamiltonian Treatment of the Spin Hamiltonian

★ In order to arrive at the effective Hamiltonian of EPR (and NMR) spectroscopy, we take the point of view

1. H0 is the Born-Oppenheimer Hamiltonian and H1 contains all spin dependent (relativistic) and all magnetic field dependent terms. 2. We assume that we know the spectrum of eigenfunctions of the BO problem (which will never be true - this makes effectiveHamiltonian theory academic in this field - progress comes from linear response to be discussed later). 3. Our model space consists of the 2S+1 functions ψSM with M=S, S-1,...,-S

(,magnetic sublevels‘) that belong to the lowest eigenvalue, E0, of the BO Hamiltonian. These functions are all degenerate within the BO approach.

4. This procedure is usually well defined: the matrix elements of H1 are much smaller than those of the BO Hamiltonian and typically the outer ,b‘ space is well removed from the ,a‘ space - only for orbitally (nearly) degenerate states (such as Jahn- Teller systems the treatment breaks down) Defining the Spin-Hamiltonian

★ Now that we can write our 0th order functions as: αSM (α = a or b)

★ We arrive at the effective Hamiltonian:

eff (0) (1) ′ th (H )MM ′ = E0 + aSM | H | aSM The 0 order ground state energy can obviously be dropped since it does aSM | H (1) | bS′M ′ bS′M ′ | H (1) | aSM ′ not add anything to the splitting of the − magnetic sublevels. ∑ (0) (0) ′ ′′ E − E bS M b 0 ★ But there is a deep symmetry that relates the components with different M for each state ,a‘ or ,b‘ - we have to make use of this with the powerful Wigner- Eckart theorem in the next step. ★ But let us first be more specific on the perturbing Hamiltonian and derive the g- Tensor. Let:

(1) li = Angular momentum of electron i relative to H = HLS + HSOC the ,global‘ origin (whatever this means ...) = (ˆ + g ˆ )+ SOCˆ si = Spin angular momentum of electron i βB∑ li e si ∑hi si hSOC = Effective one-electron spin-orbit Hamiltonian i i (e.g. SOMF) Derivation of the g-Tensor

★ First of all, the first order terms are zero since the expectation value over the purely complex operators l or hSOC vanish:

aSM | H (1) | aSM ′ = 0

★ Hence we are interested in the second-order terms - but only those terms that are linear in the magnetic field since the g-Tensor describes a linear coupling to B. This immediately gives:

( eff ) = −β Δ−1 aSM | + 2ˆ |bS′M ′′ bS′M ′′ | SOCˆ |aSM ′ H MM ′ ∑ b B∑i li si ∑i h si bS′M ′′ − β Δ−1 aSM | SOCˆ |bS′M ′′ bS′M ′′ | + 2ˆ |aSM ′ ∑ b ∑i h si B∑i li si bS′M ′′ ★ The LS matrix element reduces easily since the orbital angular momentum part is diagonal in spin and the spin angular momentum part vanishes since it is diagonal in the spatial part;

aSM | + 2ˆ |bS′M ′ = δ aSM | |bS′M ′ B∑i li si SS′ ∑i li B The Second Order g-Tensor

★ After this significant detour we can now evaluate the sums over the intermediate M-components exactly and arrive at the second-order expression for the g- Tensor ★ Let us first look at an element of the Spin-Hamiltonian:

βB g SS | S | SS = βB g S z zz z z zz ★ Now the same for our perturbation sum:

Heff = −βB Δ−1 aSS | l | bSS bSS | hSOCsˆ | aSS ( )SS z ∑ b ∑i iz ∑i z 0,i bS − βB Δ−1 aSS | hSOCsˆ | bSS bSS | l | aSS z ∑ b ∑i z 0,i ∑i iz bS ′ ★ Thus (and generalizing to all components): ★ Note: Only excited 1 −1 SOC g = − Δ aSS | l | bSS bSS | h sˆ | aSS states of the same KL ∑ b ∑i iK ∑i L 0,i S bS spin as the ground 1 state − Δ−1 aSS | hSOCsˆ | bSS bSS | l | aSS ∑ b ∑i K 0,i ∑i iL ★ Note: Only standard S bS ′ components M=S Example 3: Size Consistent-Self Consistent CI

„Deriving approximations to the many particle electron correlation problem“ Definition of the Correlation Energy

The Hartree-Fock model is characterized by:

✓ The use of a single Slater determinant which describes a system of N quasi-independent electrons (independent particle or mean field model!)

✓ The orbitals are optimized to achieve the lowest possible .

✓ The method is variational. It provides an upper bound to the exact solution of the Born- Oppenheimer hamiltonian (usually >99% of the exact nonrelativistic energy is recovered).

✓ The remaining energy error is called correlation error and arises from „instantaneous“ electron-electron interactions (as opposed to the mean-field interaction present in HF theory).

✓ Define the correlation energy as (Löwdin): Ansatz for the Many Particle Wave Function

Ψ =C Φ + C iΦa +( 1 )2 C ij Φab +( 1 )2 C ijk Φabc + ...+(n − fold exc.) 0 HF ∑ a i 2! ∑ ab ij 3! ∑ abc ijk !i#a##"###$ !####ij#a"b #####$ !###i#jk#a#b"c ######$ Singles Doubles Triples

HC = ESC Variational H = Φ | H | Φ Principle IJ I J

SIJ = ΦI | ΦJ

⎛ ˆ ˆ ⎞ ⎜ E 0 | H | S 0 | H | D 0 0 ⎟ ⎜ HF ⎟ ⎜ ⎟ ⎜ ˆ ˆ ˆ ⎟ ⎜ S | H | S 0 | H | D S | H |T 0 ⎟ ⎜ ⎟ ⎜ ⎟ The Full-CI H = ⎜ ˆ ˆ ˆ ⎟ ⎜ D | H | D D | H |T D | H |Q ⎟ Matrix ⎜ ⎟ ⎜ ⎟ ⎜ ˆ ˆ ⎟ ⎜ T | H |T T | H |Q ⎟ ⎜ ⎟ ⎜ ⎟ ⎝⎜  ⎠⎟ Size of the Full CI space

n Number of Determinants

1 400 2 35100 3 1185600 Example: 4 19191900 10 electrons 5 16581806 50 orbitals 6 806059800 (e.g. H2O) 7 2237227200 8 3460710825 9 2734388800 10 847660528 Σ 10272278170 ~ 1010

The size of the full CI matrix is HUGE even for moderately sized systems! Model System: Minimal Basis H2

For a single minimal basis H2 we found that the CID matrix was of the form:

⎛ ⎞ Δ = Ψ | Hˆ | Ψ − Ψ | Hˆ | Ψ ⎜ 0 V ⎟ D D HF HF H = ⎜ ⎟ ⎝⎜ V Δ ⎠⎟ ˆ V = Ψ0 | H | ΨD With the lowest eigenvalue: Ground state of the Double excitation minimal basis H 2 2 2 E = 1 Δ − Δ + 4V system 0 2 ( )

By contrast, for N noninteracting H2 molecules CID gave:

E = 1 Δ − Δ2 + 4NV 2 0 2 ( )

Which is not size consistent. We return to N=2 and study what is missing from CID. Refinement of the CID treatment

One step beyond CID is to include higher excitations. In the minimal basis 2x H2

model system this would be a „simultaneous pair excitation“ in which both H2‘s are put in their excited state.

Matrix-elements: Diagonal doubles

Diagonal quadruple

Doubles/ground state

Quadruple/ground state

Quadruple/doubles =Doubles/ground state! In order to solve the problem we form again the symmetry adapted linear combination of the two doubles:

The variational principle leads us then to the CI matrix (the configurations are in the order |0>, |D>, |Q>): ⎛ ⎞ ⎜ ⎟ ⎜ 0 2V 0 ⎟ ⎜ ⎟ H = ⎜ 2V Δ 2V ⎟ ⎜ ⎟ ⎜ ⎟ ⎝⎜ 0 2V 2Δ ⎠⎟

The lowest root is (without proof; look in a formula collection or use a computer algebra system):

This is twice the energy of a single H2. Thus, the inclusion of the quadruple excitation restores the size consistency!

Furthermore:

For noninteracting subsystems, the coefficients of the quadruples are exactly products of doubles coefficients! Conclusions

We had 3 key results in studying the 2xH2 problem: 1. Inclusion of the simultaneous pair excitation exactly restores the size consistency. 2. The product of the simultaneous pair excitation was exactly proportional to the square of the coefficients of the double excitations (as predicted less rigorously but more generally by perturbation theory). 3. The matrix elements of the quadruple excitation with the doubles was equal to the matrix elements of the doubles with the ground state. Both sets of determinants differ by a double substitution from each other.

Now we want to generalize these findings and restart from the full-CI equations which are written in intermediate normalization (we neglect odd excitations at the moment): Effective (Intermediate) Hamiltonian Treatment of Higher Excitations

1. H0 is the Möller-Plesset Hamiltonian (The Fock matrix) and H1 = H - H0. Thus, the complete Hamiltonian is the Born-Oppenheimer Hamiltonian. 2. This means, we do know the eigenfunctions of the 0th order Hamiltonian exactly (Slater determinants). 3. Our model space is the HF determinant 4. Our intermediate space is the space of single and double substitutions 5. The outer-space consists of the triple-, quadruple, …. substituted determinants

„Dressing“ that incorporates the effects of higher excitations Resulting equations

Working through the algebra, we find the following intermediate effective Hamiltonian:

Ψab | Hˆ | Ψ + 1 C kl Ψab | Hˆ | Ψcd +C ij (E − Δ(EPV )) =C ij E + E ij HF 4 ∑ cd ij kl ab corr ijab ab ( HF corr ) klcd   Diagonal Shift Normal CID RHS

(EPV ) 1 cd cd Δ = C Ψ | Hˆ | Φ 2 ijab 4 ∑ kl kl HF Rigorous: (SC) -CISD klcd⊂ijab

Δ(EPV ) = 1 C kl Ψcd | Hˆ | Φ Approximate: CEPA/1 ijab 4 ∑ cd kl HF klcd⊂ijab = 1 C kl kl ||cd 4 ∑ cd klcd⊂ijab ≈ 1 2 ∑ εkl kl⊂ij (EPV=„Exclusion Principle Violating“) = ( + )− ∑ εik εjk εij k Generalization: Theory

Ψ = exp T Ψ Ψ = ... ( ) = c ( ) ( ) 0 0 φ1 φN φi r ∑ µi ϕµ r Ansatz Reference determinant MOs µ MO BFs (Coester & Kuemmel) coeffs Cluster Operator T =T +T +T + ... Cluster 1 2 3 Amplitudes T = tia+a T = 1 tij a+a+a a 1 ∑ia a a i 2 4 ∑ijab ab b a i j CC wavefucntion 1 2 Ψ = (1+(T1 +T2 + ...)+ 2 (T1 +T2 + ...) + ...)Ψ0

1 2 1 2 = (1+T1 +T2 + 2T1 +T1 T2 + 2T2 + ...)Ψ0 !#######"#######$ !#######"#######$ Connected excitations disconnected excitations like CI, linear (statistically uncorrelated motion) Determination of the energy and the cluster amplitudes

−T ˆ T ECC = Ψ0 |e He | Ψ0 Gold Standard: −T T Nonlinear equation set, R = t Ψ |e Hˆe | Ψ = 0 not hard to solve; CCSD(T) K K 0 0 up to 4th power of amplitudes Summary

✓ Effective Hamiltonians are an extremely versatile and powerful tools in electronic structure theory

✓ Effective Hamiltonians can be used as powerful numerical tools.

✓ Effective Hamiltonians can be used to derive new approximations to the many body problem that are firmly grounded in fundmental physics.

✓ Effective Hamiltonians can be used to find parameterizations for groups of low-lying states.

✓ Effective Hamiltonians can be used to connect phenomenological models to fundamental physics.

✓ Effective Hamiltonians bring structure and order into complex problems

✓ Effective Hamiltonians help to create a simple language in which we can talk about complex problems The End Example 4: Ligand Field Theory Ligand Field Theory as an Effective Hamiltonian

FeO42- doped in K2SeO4

2000

1500 )

-1 Brunold, T.C.; Hauser, A.; Güdel, H.U. J.Luminescence, 1994, 59, 321-332 cm

-1 1000

3 3 3 3 ε (M T1 T1 T1 T1 LMCT 500 NOT Covered by Ligand Field Theory 0 3 3 1 1 3 1 1 1 1E 1A1 T2 T1 T2 T1 T1 E T2 A1 d-d Covered by Ligand Field Theory 10000 20000 30000 40000 Energy (cm-1) 54 Atanasov, M.; Sivalingam, K.; Ganyushin, D.; FN Struc. Bond., 2012, pp 149 Tanabe-Sugano Diagrams

d d x2-y2 dz2 hν x2-y2 dz2

dxy dxz dyz dxy dxz dyz

3+ [V(H2O)6] „Complete“ Ligand Field Theory

A complete LFT calculation in the strong field scheme proceeds as:

1. Build the one-electron matrix:

2. Build all configurations ...

3. Build all Slater determinants: ΦI(x1,...,xN)

4. Build all Configuration State functions for total spin S and Irrep Γ

5. Calculate Hamiltonian matrix elements

6. Diagonalize the ligand field Hamiltonian

→ Yields all ligand field multiplets as a function of the LFT parameters. Order them in the Tanabe-Sugano diagrams The Angular Overlap Parameterization

In the AOM the one-electron part of the ligand field is written as: L= sum over ligands hF(, , ) F (, , ) e ; ,F= sc , angular factor (symmetry!) ab ==∑∑ λλλaθϕψ L L L b θϕψ L L L, L λ σππ L λ eσ,π= Interaction parameter (ligand specific, transferrable) Two-electron part of the ligand field

−1 −1 0 2 2 24 4 dσ dσ r12 dπ dπ = dσ dσ r12 dπ ′dπ ′ = Fdd + 49 Fdd − 441 Fdd

−1 0 4 2 36 4 didi r12 didi = Fdd + 49 Fdd + 441 Fdd

−1 0 4 2 34 4 dδ dδ r12 dδ ′dδ ′ = Fdd + 49 Fdd − 441 Fdd } Racah parameters. etc.

Just two (one) parameter describes the electron-electron repulsion, 1-3 parameters for each ligand. Examples of AOM parameters

Ligand eσ (cm-1) eπ (cm-1) CN- 7530 -930 F- 7390 1690 Cl- 5540 1160 Br- 4920 830 I- 4100 670

NH3 7030 0 en 7260 0 py 5850 -670

H2O 7900 1850 OH- 8670 3000

Note: 10Dq = Δ = 3eσ-4eπ .... The task at hand is to construct an effective Hamiltonian that yields the ligand field states (and only those!)

.... This effective Hamiltonian will be identified with the complete ligand field CI matrix

.... It will turn out that this match will allow for an unambiguous determination of the ligand field parameters The ab initio Intermediate Hamiltonian

1. We have a „model space“ that contains all the essential physics that we want to describe. This is the CAS(n,5) space of N-particle wavefunctions that cleanly maps onto the ligand field manifold 2. We have a „outer space“ that brings in all the remaining effects of dynamic correlation

✓ intermediate Hamiltonian in the model space. ✓ Same dimension as ligand field CI matrix ✓ Completely ab initio. No parameters! ✓ Yields a near exact eigenvalue spectrum

If this concept is realized in the CASSCF/NEVPT2 framework we obtain the QD- NEVPT2 method

An element of CAS-CI matrix

Interaction of CAS-CSFs with NEV outer space

Energy of outer space functions using the Dyall Hamiltonian Atanasov, M.; Sivalingam, K.; Ganyushin, D.; FN Struct. Bond, 2012, 143, 149-220 Unambiguous Match between NEVPT2 and LFT

Overwhelming importance:

There is a 1:1 correspondence between the ligand field CSFs and the CAS-CI CSFs.

Ligand field pure d-orbital Ab initio molecular orbital with metal d- parentage Thus, all we have to ensure is that ligand field d-orbitals and CASSCF molecular orbitals of the same parentage are ordered in the same way and that CSFs are constructed in the same way.

The condition is then that the ligand field CI matrix should resemble the ab initio effective Hamiltonian as closely as possible

For each matrix element!

Atanasov, M.; Sivalingam, K.; Ganyushin, D.; FN Struct. Bond, 2012, 143, 149-220 While this looks at first sight to be a nonlinear optimization problem, in reality things are easy because the ligand field matrix is linear in each and every ligand field parameter!

This ensures that there is a unique least squares solution that provides the unambiguous best fit of the ligand field and effective Hamiltonian matrices:

The k‘th ligand field parameter

This implies the strategy:

1. Choose your AOM scheme 2. Perform a QD-NEVPT2 calculation to obtain Heff 3. Solve linear equation system to obtain the ligand field parameters

Atanasov, M.; Sivalingam, K.; Ganyushin, D.; FN Struct. Bond, 2012, 143, 149-220 Application to CrX63- (X=F,Cl,Br,I) Prof. Mihail Atanasov

CrF 3- CASSCF NEVPT2 Calculated spectra6 Deviations from Ligand field fit 4 -20 -111 T 2 CASSCF NEVPT2 exp and ab initio calculations 4 4T1(1)13380 -42 15365 -533 15200 T2 Extracted parameters 4 T1(1) 4T1(2)21424 -16 23449 439 21800 4T1(2) 34778 35307 35000 2E(1) -64 -283 2E(1) CrF63-19409 17579 CrCl63-16300 CrBr63- CrI63- 2T1(1) -38 -575 2T1(1)CASSCF NEVPT220430 Exp. 18681CASSCF NEVPT216300 Exp. CASSCF NEVPT2 Exp. CASSCF NEVPT2 2 -85 -119 2T2(1) T2(1)27372 25288 23000 10Dq 13359 15255 15297 10941 14109 12630 9876 13542 13089 9258 13533 2A1 2A1 29674 25 30022 -156 B 2 1071 863 32664734 32979988 767 632 972 729 471 943 699 T2(2) 2T2(2) 88 -527 2T1(2) 33736 33800 C 4018 37202T1(2) 349220 3907 -6783605 3180 3886 3631 3249 3841 3627 2E(2) 36009 35316 2E(2) 10 -151 C/B 2T1(3) 3.75 4.3139668 4.76 39324 3.95 4.70 5 4.00 4.98 6.90 4,07 5.18 2 90 -694 2T2(3) T1(3)46237 45791 ✓ Very good agreement between ab initio 2T1(4) 2T (3)46546 -24 48689 -797 2 values and empirical values for 10Dq 2A2 51121 47612 2T1(4) 89 -767 ✓ Still slight overestimation of electron 2 T2(4) 52510 49902 nd 2A2 -4 -480 repulsion (basis set incompleteness + 2 2E(3) 55140 55611 order perturbation theory) 2 -11 762 T2(4) ✓ Ligand field fit to CASSCF near perfect, to 2E(3) -166 -1017 NEVPT2 within 1000 cm-1

Atanasov, M.; Sivalingam, K.; Ganyushin, D.; FN Struct. Bond, 2012, 143, 149-220