-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy paththesis_chap5.tex
3231 lines (3024 loc) · 168 KB
/
thesis_chap5.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
% Transition metal complex/coordination: Fe3+ in mussels, Zn/Cu in spiders
% https://www.nature.com/articles/srep02914#ref2
% Mechanics of metal-catecholate complexes: The roles of coordination state and metal types
% % Perovskite in Earth’s deep interior
% % http://science.sciencemag.org/content/358/6364/734
% 1969: pressure-induced SCO~\cite{Ewald1969}.
% % work term P delta V, delta V ~ -10 to -20 A^3
% % related to changes in earth crust, mantle
% 1992: first SCO device~\cite{Kahn1992}.
% % Review on photochemistry of iron
% \cite{Chen2018}.
\chapter{Photoinduced Spin Crossover in Iron(II) Systems}
\label{ch: SCO}
Some transition-metal complexes can change spin states under certain perturbation of
their environment. This process is called spin crossover (SCO) and this chapter is focused
on its role in the light-induced dynamics of two iron(II) compounds:
[Fe\textsuperscript{II}(bpy)\textsubscript{3}](PF\textsubscript{6})\textsubscript{2}
and [Fe\textsuperscript{II}(PM-AzA)\textsubscript{2}(NCS)\textsubscript{2}],
abbreviated herein to BPY and AZA respectively.
%
The general physics of SCO will be described in some details to provide a basis to
understand the findings of the two sections that follow. Herein, I will present
the results from two experiments that studied the two aforementioned molecules
using transient absorption (TA) spectroscopy and ultrafast electron diffraction (UED).
This discussion is mostly based on two recent articles ``Spectral Signatures of Ultrafast
Spin Crossover in Single Crystal
[Fe\textsuperscript{II}(bpy)\textsubscript{3}](PF\textsubscript{6})\textsubscript{2}''
and ``Structural Dynamics upon Photoexcitation in a Spin Crossover Crystal Probed with
Femtosecond Electron Diffraction,'' previously published in the journals
Chemistry --- A European Journal and Angewandte Chemie respectively~\cite{Field2016, Jiang2017}.
% "Alternative fact" -> AzA/polypy chapters
% Real 2015, Dalton Trans. -> influence of counterion in SCO
\section{Overview of Spin Crossover}
\label{sec: SCO-overview}
% Discovery of SCO
% Theory: Pauling (Figure 8 from Pauling1932)
% Experiment: Cambi et al (Figure 2 from Cambi1933)
In 1931--1932, Linus Pauling%
\footnote{Linus Carl Pauling (1901--1994) was awarded the 1954 Nobel Prize in Chemistry
for his research on the nature of the chemical bond~\cite{Nobel1942}.} published a series of four papers
wherein the nature of chemical bonding and molecular structure is explained
in terms of quantum mechanics~\cite{Pauling1931a, Pauling1931b, Pauling1932a, Pauling1932b}.
%
In one section on the magnetic moment and electronic configuration of transition-metal complexes%
\footnote{Transition-metal complexes are chemical compounds that are formed
by various molecules attaching to a central transition-metal atom via coordination bonds;
the coordinating molecules are referred to as `ligands'~\cite{Brock1983}.}
(see Fig.~\ref{fig: SCO-overview}a),
he recognized that some metal ions can be in either of two ground states
with different number of unpaired electrons%
\footnote{In the case of iron(III) in [Fe\textsuperscript{III}L\textsubscript{6}]\textsuperscript{3+},
the five electrons in the 3d~orbitals are either all unpaired or paired twice with one left;
the former is the `high-spin' (HS) state, with total spin quantum number $S = \frac{5}{2}$;
the latter is the `low-spin' (LS) state, with $S = \frac{1}{2}$.}
depending on the bonding environment and that both
--- one `low-spin' (LS) and the other `high-spin' (HS) ---
can be present simultaneously in ratios determined by
the energy difference between them~\cite{Pauling1932a, SCO-I}.
%
This insight proved prescient when Livio Cambi and his collaborators in the early 1930s
measured the magnetic susceptibility $\chi$ of some iron(III) compounds%
\footnote{These were a homologuous series
of tris(\textit{N},\textit{N}'-dialkyldithiocarbamate)iron(III) derivatives,
(R\textsubscript{2}NCS\textsubscript{2})Fe\textsuperscript{III} where R = C$_n$H$_{2 n + 1}$.}
as a function of temperature~\cite{Cambi1931, Cambi1933}.
As seen in Fig.~\ref{fig: SCO-overview}b,
instead of a trend following the Curie-Weiss~law,%
\footnote{Named after Pierre Curie (1859--1906) and Pierre-Ernest Weiss (1865--1940),
this law describes the magnetic susceptibility $\chi$ of a material as a function of
its temperature $T$: $\chi = \chi_0 + C (T - T_\text{C})^{-\gamma}$,
where $\chi_0, C, T_\text{W}, \gamma$ are the Pauli susceptibility, Curie constant,
Weiss temperature, and critical exponent~\cite{Nobel1901, AshcroftBook}.}
they found values consistent with a LS electronic configuration at low temperature
and a HS one at high temperature and concluded correctly that
these molecules behaved as a mixture of two spin isomers
in thermal equilibrium with each other.
%
Over the next few decades, similar anomalous magnetic behaviours have been reported
in many other complexes of Fe\textsuperscript{III} and
those of Fe\textsuperscript{II}, Co\textsuperscript{II}, Co\textsuperscript{III},
Ni(II), Ni\textsuperscript{III}, Mn\textsuperscript{II}, Mn\textsuperscript{III}, Cr\textsuperscript{II},
and Cr\textsuperscript{III}~\cite{SCO-II}.
%
Concurrent theoretical developments in the form of crystal field theory~(CFT)
and ligand-field theory~(LFT)
(see Sec.~\ref{sec: SCO-theory}) allowed these observations to be
rationalized into a self-consistent physical description.
\begin{figure}[t!]
\centering
\includegraphics[width = \textwidth]{Figures/fig_SCO_overview.pdf}
\caption[Overview of spin crossover as a phenomenon.]{
Overview of spin crossover as a phenomenon.
(a) Energy level diagram by Pauling, showing the interchanging stability
of the high- and low-spin states of
a [Fe\textsuperscript{III}X\textsubscript{6}]\textsuperscript{3+} complex,
where X is a ligand, as a function of decreasing electron affinity.
(b) Magnetic susceptibility of tris(\textit{N},\textit{N}'-dialkyldithiocarbamate)iron(III)
as a function of temperature, for different alkyl chain length $n$.
(c) Magnetic Susceptibility of some Fe(II) complexes as a function of temperature,
showing abrupt transitions between LS ($S = 0$, $L = 0$) and HS ($S = 2$, $L = 2$) states,
where $\mu_{\boldsymbol{S} + \boldsymbol{L}} = \sqrt{4 S(S + 1) + L(L + 1)} \mu_\text{B}$,
$\mu_\text{B}$ is the Bohr magneton, $S$ and $L$ are the total spin and orbital quantum numbers.
(d) $^{57}$Fe~Mossba\"{u}er spectra of [Fe\textsuperscript{II}(phen)\textsubscript{2}(SCN)\textsubscript{2}]
as a function of temperature;
two distinct subspectra contribute in turn to the overall relative change in intensity,
supporting the idea of an electronic spin transition at each Fe\textsuperscript{II} site.
Panels are adapted with permission from Refs.~\cite{Pauling1932a, Cambi1933, Baker1964, Dezsi1967}
respectively.
}
\label{fig: SCO-overview}
\end{figure}
Leslie E. Orgel%
\footnote{British chemist Leslie E. Orgel~(1927--2007) is most known
for his early application of LFT to transition metal chemistry
and later proposal of the `RNA World' hypothesis for
the origin of life on Earth~\cite{Orgel1994, Joyce2007}.}
deduced from reports of divergent ground spin states of two similar Fe(II) complexes ---
HS for $\mathrm{[Fe^{II}(phen)_3]^{2+}}$ and LS for $\mathrm{[Fe^{II}(me-phen)_3]^{2+}}$,
where phen = 1,10-phenanthroline and me-phen = 2-methyl-1,10-phenanthroline ---
that the two systems lie near but on opposite sides of
the `spin crossover'~(SCO) point in the Pauling diagram
(Fig.~\ref{fig: SCO-overview}a)~\cite{Irving1953, Orgel1956}.
%
Great interest followed, as various groups attempt to map out
this SCO region as a function of metal--ligand interaction strength
by systematically varying the metal cations
and modifying the chemistry of
the ligand molecules~\cite{Figgins1960, Stoufer1961, Madeja1963, White1964, Ewald1964}.
%
In the case of Fe(II) complexes, Baker and Bobonich~\cite{Baker1964}
observed an abrupt thermal transition in the magnetic moment of
$\mathrm{[Fe^{II}(phen)_2(NCS)_2]}$, $\mathrm{[Fe^{II}(phen)_2(NCSe)_2]}$,
and $\mathrm{[Fe^{II}(bpy)_2(NCS)_2]}$
--- where bpy = 2,2'-bipyridine --- consistent with SCO (see Fig.~\ref{fig: SCO-overview}c);
however, they rejected the idea of a spin transition
and misattributed this anomaly to antiferromagnetic interaction
between two nearby $\mathrm{Fe^{II}}$ ions since they could not find any corresponding change in
their $\mathrm{^{57}Fe}$ M\"{o}ssbauer absorption%
\footnote{German physicist Rudolf Ludwig M\"{o}ssbauer (1929--2011) carried out
the first experimental observation of recoilless nuclear resonance absorption
of gamma rays in 1955--1957~\cite{Mossbauer1958};
this discovery earned him a share of the 1961 Nobel Prize in Physics~\cite{Nobel1942}.} spectra~\cite{Collins1966}.
Two years later, K\"{o}nig and Madeja~\cite{Konig1966} repeated the M\"{o}ssbauer measurements
and were able to clearly observe the appearance of a narrow doublet signal%
\footnote{The first excited state of the \textsuperscript{57}Fe nucleus has
a non-zero electric quadrupole moment which gives rise to a doublet
in its M\"{o}ssbauer absorption spectrum;
this splitting, $\Delta E_\text{Q}$, is narrowed in the LS electronic configuration
due to depopulation of metal--ligand orbitals and
thus reduced anisotropy of the electric-field gradient~\cite{Gutlich2012}.}
characteristic of a LS state upon sample cooling.
These results were soon reproduced by D\'{e}zsi et~al~\cite{Dezsi1967}
(see Fig.~\ref{fig: SCO-overview}d),
thus unambiguously establishing the existence of SCO in Fe(II) systems.
Indeed, the discovery of the M\"{o}ssbauer effect~\cite{Mossbauer1958} in 1958
proved pivotal to study of SCO
since anomalies in magnetic susceptibility $\chi_\text{M}(T)$ are not infrequently reported
and can be caused by ferromagnetic impurity in the bulk or exchange interactions
between nearby metal centres of oligomeric species~\cite{Ewald1964}.
M\"{o}ssbauer spectroscopy provides more directly well-defined information on
the oxidation and spin state of the targeted atoms.%
\footnote{Note that, for two spin states to be spectrally resolved,
it is necessary that:
(1) the relaxation time for the $\text{LS} \rightleftharpoons \text{HS}$ fluctuation is longer than
the half-life of the nuclear isomer used ($t_{1/2} = 98.3$~ns for \textsuperscript{57}Fe$^*$) and
(2) their Lamb-M\"{o}ssbauer factors $f_\text{HS}, f_\text{LS}$ are equal~\cite{SCO-I}.}
\subsection{Theoretical Considerations}
\label{sec: SCO-theory}
A body of work has been developed following Pauling
to quantitatively describe the chemical and physical properties
of transition-metal complexes~\cite{Griffith1957, FiggisBook}.
%
In 1929, Hans A. Bethe%
\footnote{See Fn.~\ref{fn: MottBethe}.} and others~\cite{Bethe1929, Vleck1932, Schlapp1932}
sought to explain the spectral lines of metallic compounds
by considering the effect of an external electrostatic field of known symmetry
on atomic energy levels; his solution is to expand the perturbing potential into its multipole components,
apply first-order perturbation theory, and take a group theoretic approach
to evaluate the necessary matrix elements.
%
The result is crystal field theory~(CFT) wherein the energy degeneracy of some atomic orbitals,
such as the d~orbitals of transition-metal atoms,
can be lifted by the anisotropic electrostatic field generated by neighbouring atoms;
an energy gap of width~$\Delta$ is opened as the orbitals are subject to
different amount of coupling depending on their spatial extent
and the symmetry of the interaction.
\begin{figure}[t!]
\centering
\includegraphics[width = \textwidth]{Figures/fig_SCO_theory.pdf}
\caption[Metal--ligand interaction for a d$^6$ electron system.]{
Metal--ligand interaction for a d$^6$ electron system.
(a) Illustration of the spatial distribution
of the five 3d~orbitals for a hydrogen-like atom coloured by phase
and grouped by symmetry --- E\textsubscript{g} and T\textsubscript{2g}.
(b) Schematic of the electrostatic interactions in crystal field theory
that lead to the d--d energy splitting $\Delta_\text{oct}$;
(c) Molecular orbital diagram of a [ML\textsubscript{6}]\textsuperscript{n+} complex,
showing the energy, electron occupancy, and symmetry of the relevant HOMOs and LUMOs;
inside the boxed area, the ligand-field MOs derived from three types of ligand are shown:
$\unslant[-.2]\sigma$~donor (left), $\unslant[-.2]\pi$~donor (centre), $\unslant[-.2]\pi$~acceptor (right);
the d~electrons of the cation is coloured in red;
those from the $\unslant[-.2]\sigma$ and $\unslant[-.2]\pi$~orbitals of the ligands are coloured in blue and green respectively.
Panel~c is adapted from Ref.~\cite{FiggisBook}
and the MOs herein is labeled by their Mulliken term symbols (see App.~\ref{ap: term-symbols}).
}
\label{fig: SCO-theory}
\end{figure}
For first-row transition metals like iron,
the relevant orbitals are the five 3d~orbitals --- those with quantum numbers $n = 3, \ell = 2$ ---
and they are shown in Fig.~\ref{fig: SCO-theory}a for reference.
%
In free space, the associated energy levels are degenerate.
When the metal cation is surrounded by coordinating ligand anions, perturbation occurs in two steps:
all the levels are first equally raised due to repulsion with the radial component of this ligand field;
they are then differentiated by the azimuthal component.
%
In Fig.~\ref{fig: SCO-theory}b, this process is illustrated for
a metal--ligand coordination with octahedral symmetry,
wherein the ligand electrons approaching along the x, y, z axes repel
electrons occupying the d$_{x^2 - y^2}$, d$_{z^2}$ orbitals more than
those of the d$_{xy}$, d$_{yz}$, d$_{zx}$ orbitals.
%
This results in different CFT energy splittings
which can be quantified in multiples of $Dq$,
a notation due to Bethe in Ref.~\cite{Bethe1929}
where $D$ is the scaling factor of the $l = 4$~component of the perturbing potential
and $q$ contains the integral over the orbitals.
%
In particular, d$_{x^2 - y^2}$ and d$_{z^2}$ are raised energetically by $6Dq$
while the other three are lowered by $4Dq$,
leading to a splitting of $\Delta_\text{oct} = 10Dq$.
%
Since values of $\Delta$ happen to lie in the visible region of the EM spectrum
(see Tables~\ref{tab: cft} and~\ref{tab: cft-full}),
light absorption across this d--d energy gap gives rise to
some of the striking ligand-dependent colours
that characterize transition-metal complexes.
% Fill the orbitals
To find the ground electronic configuration of the metal cation under this crystal field,
the d~electrons added sequentially according to Hund's rules%
\footnote{Friedrich Hund~(1896--1997) formulated three simple empirical rules to determine
the ground electronic state of an atom: maximize $S$ and $L$ while minimizing $J$,
where $S, L, J$ are the total spin, orbital, and net angular momentum numbers~\cite{Hund1925}.
\label{fn: Hund}},
filling up the lowest unoccupied orbitals first before reaching double occupancy.
%
In this scheme, transition metal cations with four to seven d~electrons ---
referred as d$^4$, d$^5$, d$^6$, and d$^7$ electron systems ---
can have two possible electron configurations:
a low-spin one where electrons are paired in the lower set of orbitals
and a high-spin one where they are left unpaired and spread maximally~\cite{Griffith1957}.
%
Generally, the HS configuration is favoured if the perturbing field is weak and
the energy splitting~$\Delta$ is small.
In the strong-field case, the LS one may be realized in the ground state.
%
More precisely, the energy gain from placing one or two electrons in the lower three orbitals
needs to be considered in opposition to
the energy of increased electron--electron Coulomb repulsion~$\Pi_\text{c}$ from double occupancy
and that of exchange interaction~$\Pi_\text{e}$ from spin pairing.
%
Therefore, it is expected that the ground state of a transition-metal complex
is HS if $\Delta < \Pi$ and LS if $\Delta < \Pi$, where $\Pi = \Pi_\text{c} + \Pi_\text{e}$
is the mean spin-pairing energy (see Fig.~\ref{fig: SCO-theory}b).
%
In Tab.~\ref{tab: cft}, values of $\Delta$ and $\Pi$ are listed for a few octahedral complexes
and it can be seen that $\Delta \sim \Pi$ may be achievable
by chemically substituting and modifying the coordinating ligands.
%
\begin{table}[ht!]
\centering
{\renewcommand*{\arraystretch}{1.5}
\begin{tabular}{ c c c c c c }
\toprule
\multirow{2}{*}{
\begin{minipage}[c]{1.75cm} \centering Electron System \end{minipage}
} & \multirow{2}{*}{Ion}
& \multicolumn{3}{c }{$\Delta_\text{oct}$ ($\times 10^3$~cm$^{-1}$)}
& \multirow{2}{*}{
\begin{minipage}[c]{2.25cm} \centering $\Pi$ ($\times 10^3$~cm$^{-1}$) \end{minipage}
} \\
\cline{3-5}
& & Cl$^-$ & OH$_2$ & CN$^-$ & \\
\midrule
d$^4$ & Cr$^{2+}$ & 10 & 13 & & 23.5 \\
& Mn$^{3+}$ & 17.5 & 20 & 31 & 28 \\
d$^5$ & Mn$^{2+}$ & 7.5 & 8.5 & 33 & 25.5 \\
& Fe$^{3+}$ & & 14 & 35 & 30 \\
d$^6$ & Fe$^{2+}$ & & 10 & 32 & 17.6 \\
d$^7$ & Co$^{2+}$ & 7.7 & 9.3 & & 22.5 \\
\bottomrule
\end{tabular}
}
\caption{Octahedral CFT splitting energy~$\Delta_\text{oct}$ and
mean spin-pairing energy~$\Pi$ of some transition-metal complexes
of the form [ML$_6$]$^{n+}$, where $\text{L} \in \{ \text{Cl}^-, \text{OH}_2, \text{CN}^- \}$.
Data is taken from Refs.~\cite{FiggisBook, Griffith1957}.
An extended version of this table is shown as Tab.~\ref{tab: cft-full} in App.~\ref{ap: cft},
where values for other metal cations and ligands are included for comparison.
}
\label{tab: cft}
\end{table}
% Problems of CFT
Indeed, crystal field theory is a conceptually simple and useful description of metal--ligand interactions.
However, a closer inspection of experimental data would show that it is inadequate.
An obvious example is demonstrated by the `spectrochemical series,' an empirical list of ligands
sorted by the value of $\Delta$ that they generate in different transition-metal complexes.%
\footnote{Japanese chemist Tsuchida Ryutaro (1903--1962) measured the absorption spectra of
several transition-metal complexes and concluded that ligand substitution
can modulate the energy of electronic transitions~\cite{Ryutaro1938a, Ryutaro1938b};
the approximate sequence at which this occurs in increasing order is known as
the spectrochemical series:
I$^-$, Br$^-$, SCN$^-$, Cl$^-$, NO$_3^-$, N$_3^-$, F$^-$, OH$^-$, OH$_2$, NCS$^-$, NH$_3$,
py, en, bpy, phen, NO$_2^-$, PPh$_3$, CN$^-$, CO, NO$^+$~\cite{FiggisBook}.}
As seen in Tab.~\ref{tab: cft}
and more extensively in Tab.~\ref{tab: cft-full} in App.~\ref{ap: cft},
neutral molecules such as OH\textsubscript{2} and NH\textsubscript{3} generate larger values of $\Delta_\text{oct}$ than any halide anions,
an inexplicable feat if only electrostatic interactions \textit{\`{a} la} CFT are considered.
% Ligand-field theory
A remedy to the limitations of CFT is to extend the scope of the metal--ligand interaction physics
from pure electrostatics to a fuller quantum-mechanical treatment
where electrons can be transferred and covalent bonds form~\cite{Griffith1957}.
%
In 1935, John H. van Vleck%
\footnote{John H. van Vleck (1899--1980) was awarded the 1977 Nobel Prize in Physics,
along with Philip W. Anderson and Nevill F. Mott (see Fn.~\ref{fn: MottBethe}),
for his work on the electronic structure of magnetic materials~\cite{Nobel1971}.} did exactly that
by applying the Hund-Mulliken molecular orbital method%
\footnote{Friedrich Hund~(see Fn.~\ref{fn: Hund}), Robert S. Mulliken~(1896--1986),
and others calculated the electron wavefunctions of a molecule as
a linear combination of those of the constituent atoms;
Mulliken was awarded the 1966 Nobel Prize in Chemistry for his contribution~\cite{Nobel1972}.}
to resolve bonding in transition-metal complexes,
thus introducing ligand-field theory (LFT)~\cite{Vleck1935}.
% Describe MO diagram
As described in Ref.~\cite{FiggisBook} and illustrated in Fig.~\ref{fig: SCO-theory}c,
LFT involves collecting together the frontier orbitals of the metal cation
--- 3d, 4s, and 4p for a first-row transition metal like Fe$^{2+}$ ---
and those of the ligands --- $\unslant[-.2]\sigma$, $\unslant[-.2]\pi$, or $\unslant[-.2]\pi^*$ ---
and sorting them by symmetry.
Those with the same irreducible representation ---
a\textsubscript{1g}, e\textsubscript{g}, t\textsubscript{2g}, etc. ---
would `interact' and be linearly combined into a new set of ligand-field molecular orbitals (MOs).
MOs constructed from in-phase linear combinations are known as bonding orbitals
and are energetically favoured since they shield the positive nuclei of the metal and ligands,
binding them together.
On the other hand, out-of-phase MOs are anti-bonding and high-energy
since they are spatially distributed everywhere else but between the interacting nuclei;
unaffected orbitals are simply non-bonding.%
\footnote{The anti-or non-bonding nature of an MO is denoted by an `n' or asterisk
to the right of its term symbol in superscript.}
In particular, the frontier electronic configuration of the LS and HS states involved in d$^6$~SCO
are recovered as (t$_\text{2g}$)$^6$
and (t$_\text{2g}$)$^4$(e$_\text{g}^{*}$)$^2$ respectively;
their corresponding term symbols are then $^1$A$_\text{1g}$ and $^5$T$_\text{2g}$~\cite{SCO-I}.
% Discussion of LFT results
From this MO-based description of metal--ligand interactions,
some key features of SCO can be understood.
% 1. possible transitions in spectra
First, the characteristic absorption bands of transition-metal complexes amenable to SCO
can be classified amongst four possible types of electronic transitions:
a X--Y charge transfer (XYCT), where X,~Y = metal or ligand.
A metal--ligand charge transfer~(MLCT) involves excitation from
a filled metal-derived MO ($\mathrm{t_{2g}}$) to an empty ligand-derived one ($\unslant[-.2]\pi^*$);
conversely, a ligand--metal charge transfer~(LMCT) is a transition from
either ligand $\unslant[-.2]\sigma$, $\unslant[-.2]\pi$ MOs to the empty metal $\mathrm{e_g^*}$ MOs.
Ligand--ligand and metal--metal charge transfers (LLCT and MMCT)
are simply ligand- and metal-centered transitions,
e.g.~$\unslant[-.2]\pi \rightarrow \unslant[-.2]\pi^*$ and
$\mathrm{t_{2g}} \rightarrow \mathrm{e_g^*}$ respectively.
% 2. ligand bonding type -> order in delta
Second, the variation in the value of $\Delta$ for different ligands,
as empirically expressed by increasing order in Ryutaro's spectrochemical series,
can be explained by the presence or absence of ligand MOs with symmetries compatible with
the $d$~orbitals of the metal cation.
%
A $\unslant[-.2]\pi$~donating ligand like any halide anion has filled $\unslant[-.2]\pi$~orbitals
from which many electrons can be transferred to overfill the ligand-field MOs,
leaving only a small HOMO--LUMO gap.
%
On the other hand, a $\unslant[-.2]\pi$~acceptor like CN$^-$ and 2,2'-bipyridine~(bpy)
have empty $\unslant[-.2]\pi^*$~orbitals which additionally stabilize
the metal-centered t\textsubscript{2g}~orbitals while leaving them available
for the metal valence electrons to occupy, creating a larger HOMO--LUMO gap
than what a plain $\unslant[-.2]\sigma$-donating ligand like OH$_2$ would generate.
% 3. assignment of upper ligand-field e_g to anti-bonding orbital -> bond elongation
Third, X-ray crystallography long revealed that the cation--anion interatomic distances
amongst first-row transition metal compounds does not decrease monotonically
as a function of atomic number; instead, there is a dependence on
the electronic spin state of the metal cation~\cite{Santen1952, Shannon1969}.
%
By inspection of the MO diagram in Fig.~\ref{fig: SCO-theory}c,
this behaviour can be rationalized by the assignment of
the upper ligand-field $\mathrm{e_g}$ energy levels to anti-bonding orbitals.
Conversion to the HS state requires occupation of $\mathrm{e_g^*}$
and lead to longer bond lengths between the metal cation and the ligands.%
\footnote{For a typical $\mathrm{Fe^{II}N_6}$-type SCO complex,
$r_\text{LS} \approx 2.0$~\AA{} and $r_\text{HS} \approx 2.2$~\AA{}~\cite{Shannon1969, Orpen1989}.}
% 4. SCO is not just delta = P
Fourth, SCO does not occur simply when the ligand-field energy splitting~$\Delta$
is roughly equal to the mean spin-pairing energy~$\Pi$ or,
more specifically, $| \Delta - \Pi | \sim k_\text{B} T$ for some temperature~$T$.
Such a condition is demonstrably unphysical since it assumes \textit{\`{a} la} Franck-Condon
that there is a single fixed value of $\Delta$ during SCO,
i.e.~vertical transition between states according to the Tanabe-Sugano term diagram%
\footnote{Yukito Tanabe (1927--present) and Satoru Sugano (1928--present)
calculated the electronic ground- and excited-state energy of
all possible electronic configurations of the 3d~orbitals
and plotted them in diagrams that clearly explain the absorption spectra of
various transition-metal complexes~\cite{Tanabe1954a, Tanabe1954b, Tanabe1956}.}
(Fig.~\ref{fig: SCO-tanabe}a).
%
A corollary of the previous discussion is $r_\text{HS}/r_\text{LS} > 1$;
considering that $\Delta_\text{X} \propto \frac{1}{r_\text{X}^5}$
where $r$ is the metal--ligand distance,
the LS and HS states have significantly different values of $\Delta$.
%
In fact, this subtlety is essential to the process of SCO:
proximity of $\Delta_\text{LS}$ to $\Pi$ allows thermal population of the higher-energy HS state
which, upon relaxation of the metal--ligand bond, can have such a reduced energy splitting
that it is stabilized against spin reversion.
%
Therefore, as first noted in Ref.~\cite{Ewald1964},
a more likely and mechanistic predictor on the occurrence of SCO
is $\Delta_\text{HS} < \Pi < \Delta_\text{LS}$
such that $\Delta E_\text{HL}^{(0)} \sim k_\text{B}T$,
where $\Delta E_\text{HL}^{(0)}$ is the zero-point energy difference between
the adiabatic molecular potentials of the HS and LS states
at their respective equilibrium nuclear configuration (Fig.~\ref{fig: SCO-tanabe}b).
\begin{figure}[t!]
\centering
\includegraphics[width = \textwidth]{Figures/fig_SCO_tanabe.pdf}
\caption[Mechanism of SCO.]{
Mechanism of SCO.
(a) Tanabe-Sugano diagram for the d$^6$~electron system,
showing the energy of various electronic states relative to
the ground state as a function of ligand-field splitting~$\Delta$;
$E, \Delta$ are normalized to the Racah parameter~$B$
and relevant states are labelled by their Mulliken term symbol.
The right panel shows the UV--Vis~absorption spectrum of
single-crystal [Fe\textsuperscript{II}(ptz)$_6$](BF$_4$)$_2$
in the LS and HS states ($T_\text{c} = $~K, ptz = 1-propyl-1\textit{H}-tetrazole);
the $\gtrsim 25000$-cm$^{-1}$ absorption shoulder is part of
metal--ligand charge transfer~(MLCT) bands that involve excitations to
the empty $\unslant[-.2]\pi^*$~MOs of the ligands (see Fig.~\ref{fig: SCO-theory}c).
(b) Schematic of the adiabatic molecular potential of
the LS and HS states as a function of metal--ligand bond length
and under different ligand-field conditions;
vibrational energy levels are drawn to highlight
the higher density of states of the HS state.
(c) Gibbs free energy~$G$ as a function of high-spin fraction~$x_\text{HS}$
and temperature~$T$ relative to $T_\text{c}$,
calculated from a Ising-type SCO interaction model.
(d) Thermal SCO of Fe\textsuperscript{II}(PM-AzA)$_2$(NCS)$_2$
under increasing external pressure.
Adapted respectively from
Refs.~\cite{Tanabe1954b, Marino2014, Ewald1964, Zimmermann1983, Ksenofontov1998}.
}
\label{fig: SCO-tanabe}
\end{figure}
% LS is QM ground state, HS is thermodynamic ground state for T > T1/2
% HS is entropically favoured
An addendum to the LFT-based description of SCO is
the recognition that it is a thermal phase transition,
wherein there is a complete transfer of population from
the LS state to the HS state despite $\Delta E_\text{HL}^{(0)} > 0$.
%
In the work of Sorai and Seki~\cite{Sorai1972, Sorai1974},
the mechanism of this process is elucidated
by directly studying the thermodynamics of the prototypical SCO system,
[Fe\textsuperscript{II}(phen)\textsubscript{2}(NCX)\textsubscript{2}] for X = S and Se.
%
It was found that SCO is an entropy-driven phenomenon
which occurs when the enthalpy difference caused by
the change in ligand-field energy~$\Delta H \sim \Delta E_\text{HL}^{(0)}$
is balanced out by the entropy term~$T \Delta S$ at $T = T_{1/2}$,
as shown by the application of the van't~Hoff equation,%
\footnote{Dutch physical chemist Jacobus H. van't~Hoff (1852--1911) was awarded in 1901
the first Nobel prize in chemistry for his contributions to chemical thermodynamics,
one of which is this formula on the temperature dependence of
the equilibrium constant of a simple mixture of chemical species~\cite{Nobel1901}.},
with Gibbs%
\footnote{Josiah Willard Gibbs (1839--1903) was an American mathematical physicist
whose contributions include statistical mechanics and, in particular,
the concept of `available energy' at constant volume and temperature~\cite{GibbsMemoir}.} free
energy $G = \sum_{i = 1}^{m} x_i G_i^\circ - T \Delta S_\text{mix}$
and entropy of mixing $\Delta S_\text{mix} = N k_\text{B} \sum_{i = 1}^{m} x_i \ln(x_i)$,
in chemical equilibrium, $\left( \partial G / \partial x_i \right)_{T, P} = 0 \enskip \forall i$,
%
\begin{equation}
\begin{aligned}
\ln K & = - \frac{\Delta G^\circ}{N k_\text{B} T}
& = -\frac{\Delta H^\circ}{N k_\text{B} T} + \frac{\Delta S^\circ}{N k_\text{B}}
\label{eq: vant-Hoff1}
\end{aligned}
\end{equation}
%
which becomes at the transition, $x_\text{HS} = \frac{1}{2}$ and $T = T_\text{1/2}$,
%
\begin{equation}
\begin{aligned}
T_\text{c} = \frac{\Delta H^\circ}{\Delta S^\circ}
\label{eq: vant-Hoff2}
\end{aligned}
\end{equation}
%
where $x_\text{HS}$ is the HS mole fraction,
$K = \frac{x_\text{HS}}{1 - x_\text{HS}}$ is the equilibrium constant,
$\Delta G^\circ$, $\Delta H^\circ$, $\Delta S^\circ$ are the standard Gibbs free energy,
enthalpy, and entropy difference between the HS and LS states,
$N$ is the total number of molecules;
more details can be found in Refs.~\cite{Slichter1972, Gutlich1979}.
%
Crucially, the entropy of transition~$\Delta S$ was measured to be ca.~3.6~times
larger than the expected entropy gain in the electronic degree of freedom
from the increased spin multiplicity of the HS state.
%
Instead, as suggested in Refs.~\cite{Ewald1964, Ewald1969},
it is the phonon system that provides the necessary entropy:
occupation of anti-bonding orbitals in the HS state softens
the metal--ligand potential and drives a drastic increase
in the number of thermally accessible vibrational states
(centre panel in Fig.~\ref{fig: SCO-tanabe}b)
%
Therefore, during SCO, the LS state remains the quantum-mechanical ground state;
however, as seen from the shape of $G(x_\text{HS}, T)$ in Fig.~\ref{fig: SCO-tanabe}c,
it is simply overtaken by the entropically favoured HS state
which becomes the thermodynamic ground state at high enough temperature
thanks to strong electron-phonon coupling~\cite{SCO-III}.
%
In this view, it is not unexpected that the LS--HS equilibrium of SCO
can be additionally controlled by applying external pressure, as in Fig.~\ref{fig: SCO-tanabe}d;
through the work term~$P \Delta V_\text{HL}$ and the Clapeyron relation
$\frac{d T_\text{c}}{d P} = \frac{\Delta V_\text{HL}}{\Delta S_\text{HL}}$,
where $\Delta V_\text{HL}$ is the change in molecular volume%
\footnote{From temperature-dependent XRD measurements of several Fe(II) SCO systems,
the $\text{LS} \rightleftharpoons \text{HS}$ volume contraction associated with the spin transition
is about $1.5$--$2.5$~\%~\cite{Guionneau1999, SCO-II}.\label{fn: volume-contraction}}
associated with the change in metal--ligand bond length~$\Delta r_\text{HL}$,
$\Delta E_\text{HL}^{(0)}$ increases and
the LS becomes more favoured~\cite{Ewald1969, Ksenofontov1998, Boillot2002}.
\subsection{Light-Induced Excited Spin-State Trapping}
\label{sec: SCO-photo-1}
Up to now, spin crossover has been understood as a mixture of two distinct spin isomers
in chemical equilibrium that interconverted within the ca.~100-ns
characteristic time window of $^{57}$Fe~M\"{o}ssbauer spectroscopy.
%
To refine this time scale, Beattie et~al~\cite{Beattie1973}
used laser Raman temperature jump%
\footnote{Laser Raman temperature jump is a technique used
for studying fast chemical reactions in solution;
it involves an impulsive heating of the spectating solvent
by a ns~near-IR pulse from a Q-switched neodymium-doped glass laser,
Raman-shifted from $1.41$-$\unslant\mu$m to $1.41$-$\unslant\mu$m
to match the wavelength of the first overtone of the O--H stretch~\cite{Turner1972}.}~(LRTJ)
to impulsively displace the $\text{LS} \rightleftharpoons \text{HS}$ equilibrium of
Fe\textsuperscript{II}(HB(pz)$_3$)$_2$ in solution to the right
and followed the $32$-ns relaxation of the resulting HS population,
making the first time-resolved pump-probe study of SCO.
%
% McGarvey discovery of light-driven SCO
It is in this laser-driven vein that, soon after,
McGarvey and Lawthers pumped several other Fe(II) SCO complexes%
\footnote{Fe\textsuperscript{II}(biz)$_3^{2+}$, Fe\textsuperscript{II}(ppa)$_2^{2+}$,
and Fe\textsuperscript{II}(pyimH)$_3^{2+}$ in water and acetonitrile--methanol,
where biz = 2,2'-bi-1,4,5,6-tetrahydropyrimidine, ppa = \textit{N}$^2$-(2-pyridylmethyl)picolinamidine,
and pyimH = 2-(2-pyridylimidazole)~\cite{McGarvey1982}.} in solution
with few-ns 530-nm laser pulses centered on their MLCT~bands
and observed ground state bleach~(GSB) and excited state absorption~(ESA)
in the transient absorption spectra~\cite{McGarvey1982}.
Since the decay time of these time-resolved features agrees very well
with the HS relaxation time measured using LRTJ~\cite{Reeder1978, Dose1978}
and the spectral profile of the ESA matches that of the transient HS state
in a similar non-SCO Fe\textsuperscript{II}N$_6$ complex~\cite{Creutz1980},
they proposed a theretofore-unknown spin crossover process that is
photophysically induced.
%
As seen on the right side of Fig.~\ref{fig: SCO-LIESST}c,
this spectral assignment proved correct when such a transient HS state was found~\cite{Deisenroth1994}
using $^{57}$Fe time-differential M\"{o}ssbauer emission spectoscopy~(TDMES),
a time-resolved technique that can directly probe the spin state
of iron coordination complexes~\cite{Kajcsos1986, Alflen1989, SCO-II}.
\begin{figure}[t!]
\centering
\includegraphics[width = \textwidth]{Figures/fig_SCO_LIESST.pdf}
\caption[Light-induced excited state trapping (LIESST).]{
Light-induced excited state trapping (LIESST).
Steady-state UV--Vis absorption spectra of
single-crystal [Fe\textsuperscript{II}(ptz)$_6$](BF$_4$)$_2$:
(a) during thermal spin crossover, LIESST, and reverse LIESST
(insets: photos of the sample under the different illumination condition);
(b) as a function of excitation wavelength~$\lambda_\text{exc}$.
(c) Left panel: Arrhenius plot for the $\text{HS} \rightarrow \text{LS}$ relaxation rate~$k_\text{HL}$
in single-crystal [Mn$_{1-x}^\text{II}$Fe$_x^\text{II}$(bpy)$_3$](PF$_6$)$_2$
($x = 5 \times 10^{-4}$) following MLCT excitation ($\lambda_\text{exc} = 532$~nm),
as measured by optical transient absorption and time-differential
M\"{o}ssbauer emission spectroscopy;
right panel: measured~(markers) and theoretical~(dashed line)
low-temperature $\text{HS} \rightarrow \text{LS}$ tunnelling rate~$k_\text{HL}(T \rightarrow 0)$
as a function of the SCO transition
temperature~$T_{1/2} = \Delta E_\text{HL}^{(0)} / \Delta S_\text{HL}^{(0)}$,
for a series of Fe\textsuperscript{II} SCO complexes.
(d) Schematic illustrating the photophysical mechanism of LIESST
and reverse LIESST with and without low-lying MLCT states,
as currently understood.
Panels~(a) and (b) are adapted from Ref.~\cite{Hauser1991a};
insets of Panel~(a), Ref.~\cite{Gutlich2012};
Panel~(c), Refs.~\cite{Deisenroth1994, Hauser1991c};
Panel~(d), Ref.~\cite{Hauser2017}.
}
\label{fig: SCO-LIESST}
\end{figure}
% Discovery of LIESST
In 1984, Decurtins et~al~\cite{Decurtins1984, Decurtins1985}
observed a similar non-thermal spin transition in single-crystal
$\mathrm{[Fe^{II}(ptz)_6](BF_4)_2}$, where ptz = 1-propyl-1\textit{H}-tetrazole,
by pumping instead with CW light at the ligand-field absorption bands of the LS state:
$\mathrm{^1A_{1g}} \rightarrow \mathrm{^1T_{1g}}$ ($514$~nm) and
$\mathrm{^1A_{1g}} \rightarrow \mathrm{^3T_{1g}}$ ($980$~nm).
%
Particularly, they discovered that the $\text{HS} \rightarrow \text{LS}$ relaxation
can be slowed down so much at cryogenic temperatures ($T < 50$~K)
that the molecule is effectively trapped in the HS state.
%
This phenomenon of `light-induced excited state trapping'~(LIESST)
can be clearly seen in Fig.~\ref{fig: SCO-LIESST}a,
where, under constant illumination of light of different wavelengths,
the absorption peaks associated with the LS state vanishes
and are replaced by those of the HS state;
on the macroscopic scale, this is manifested by the colour reversal
of the crystal within the beam spot, from purple back to colourless (see inset).
%
Two years later, Hauser et~al~\cite{Hauser1986a, Hauser1986b} found that
subsequent pumping into the $\mathrm{^5T_{2g}} \rightarrow \mathrm{^5E_{2g}}$
absorption band ($820$~nm) of the same compound
transfers population from the metastable HS state back to the ground LS state,
resulting in a `reverse LIESST'~(rLIESST).%
\footnote{Gall\'{e} et~al~\cite{Galle2013} showed that
photoexcitation to the $\mathrm{^5MLCT}$ band can also lead to rLIESST.}
%
This effect is readily apparent in Fig.~\ref{fig: SCO-LIESST}b
which shows the absorption spectra of a sample in the metastable HS state
after photoexcitation as a function of the excitation wavelength~$\lambda_\text{exc}$.
Indeed, pumping at the HS spectral feature around ca.~820~nm (blue) leads
to its direct replacement by an absorption band at $\lambda_\text{probe} \sim 18500$~cm$^{-1}$ (red)
that is characteristic of the $\mathrm{^1A_{1g}} \rightarrow \mathrm{^1T_{1g}}$
transition of the LS state.
% SCO forbidden under selection rules
Formally, transfer of population between states of different spin multiplicities ---
$\mathrm{^1A_{1g}}$ and $\mathrm{^5T_{2g}}$ --- ought not to happen
since such transitions are highly forbidden by the spin selection rule of electronic spectroscopy.%
\footnote{Electronic transitions with $\Delta S \neq 0$ (and $\Delta L = 0$) are prima~facie forbidden
since the corresponding transition moment integrals are identically zero by symmetry~\cite{Harris1978}.}
%
In light of the discovery of forward and reverse light-induced SCO,
models were developed to tentatively describe the physics that occurs during the photocycle.
% HS -> LS relaxation model
The first segment of the photoinduced SCO reaction to be well-understood is also
its last and slowest, thus most easily observed, segment:
the non-radiative $\Delta S = 2$ relaxation from the $\mathrm{^5T_{2g}}$ HS state
back to $\mathrm{^1A_{1g}}$ LS ground state.
%
Buhks et~al~\cite{Buhks1980} proposed that it can be described in terms of
a non-adiabatic multiphonon process occurring between two distinct spin states
--- characterized by different equilibrium nuclear configurations, separated by an energy barrier
which is large relative to $k_\text{B} T$, and coupled by spin--orbit interaction.
%
In this model (see App.~\ref{ap: sco} for more details),
assuming Fermi's Golden~Rule~\cite{Dirac1927, Fermi1950}
and transitions between a LS harmonic potential and
a HS one that is displaced energetically by $\Delta E_\text{HL}^{(0)}$
and horizontally by $\Delta Q_\text{HL} = \sqrt{6} \Delta r_\text{HL}$ along
the totally symmetric metal--ligand stretch mode,
the $\text{HS} \rightarrow \text{LS}$ relaxation rate is given by,
%
\begin{equation}
\begin{aligned}
k_\text{HL}(T)
& = \frac{2 \unslant[-.2]\pi}{\hbar^2 \omega}
| \langle \psi_\text{HS} | \hat{H}_\text{SO} | \psi_\text{LS} \rangle|^2 \bar{F}_n(T)
\end{aligned}
\label{eq: kHL}
\end{equation}
%
where $\omega$ is the frequency of the active vibrational mode,
$n = \frac{\Delta E_\text{HL}^{(0)}}{\hbar}$ is the reduced HS--LS zero-point energy gap,
$\langle \psi_\text{HS} | \hat{H}_\text{SO} | \psi_\text{LS} \rangle$ is the spin--orbit coupling matrix element,
and $\bar{F}_n(T)$ is the ensemble-averaged Franck-Condon factor.
%
% High T limit
In the high-temperature limit, $k_\text{B} T \gg \hbar \omega$,
a classical Arrhenius rate equation%
\footnote{Through this formula~\cite{Arrhenius1889},
Swedish physicist Svante A. Arrhenius (1859--1927) related
the temperature dependence of the rate of a chemical reaction
to the concepts of activation energy and Boltzmann distribution;
for related work, he was awarded the 1903 Nobel Prize in Chemistry~\cite{Nobel1901}.} is recovered,
%
\begin{equation}
\begin{aligned}
k_\text{HL} & \propto \text{e}^{-E_\text{a} / k_\text{B} T} \\
\ln(k_\text{HL}) & \approx -\frac{E_\text{a}}{k_\text{B} T}
\label{eq: kHL-highT}
\end{aligned}
\end{equation}
%
where $E_\text{a} = \frac{1}{4} \hbar \omega S (1 - \frac{n}{S})^2$
is the classical activation energy that trapped HS molecules
need to thermally overcome to reach the LS state;
this regime is manifested in the form of the sloped trendline
in the Arrhenius plot on the left side of Fig.~\ref{fig: SCO-LIESST}c.
%
% Low T limit
In the low-temperature limit, $T \rightarrow 0$,
where only the $m = 0$ vibrational level of the HS state is occupied
and the Franck-Condon factor $\bar{F}_n(0)$ is simply given by
$|\langle \chi_n | \chi_0 \rangle |^2 = \frac{S^n \text{e}^{-S}}{n!}$,
a constant $\text{HS} \rightarrow \text{LS}$ relaxation rate is obtained,
%
\begin{equation}
\begin{aligned}
k_\text{HL}
& = \frac{2 \unslant[-.2]\pi}{\hbar^2 \omega}
|\langle \psi_\text{HS} | \hat{H}_\text{SO} |\psi_\text{HS} \rangle |^2 \frac{S^n \text{e}^{-S}}{n!}
\label{eq: kHL-lowT}
\end{aligned}
\end{equation}
%
where $S = \frac{k {\Delta Q_\text{HL}}^2}{2 \hbar \omega}$
is the Huang-Rhys parameter,%
\footnote{This quantity is named after
physicist-couple Kun~Huang (1919--2005) and Avril~Rhys (1928--present),
who first introduced it in their seminal work~\cite{HuangRhys1950}.}
a dimensionless measure of the horizontal displacement of the two potential wells
and $k$ is the force constant of the metal--ligand bond.
In this regime (horizontal asymptote on the left side of Fig.~\ref{fig: SCO-LIESST}c),
the HS population can only traverse the potential barrier and
slowly return to the LS state by quantum-mechanical tunneling~\cite{Xie1987}.
%
On the right side of Fig.~\ref{fig: SCO-LIESST}c,
Eq.~\eqref{eq: kHL-lowT} is plotted using
known empirical values of the model parameters%
\footnote{From Ref.~\cite{SCO-II},
$|\langle \psi_\text{HS} | \hat{H}_\text{SO} |\psi_\text{HS} \rangle| \approx 150$~cm$^{-1}$,
$\hbar \omega \approx 250$~cm$^{-1}$, $\Delta r_\text{HL} \approx 0.2$~\AA{},
$k \approx 200$~N/m, $S \approx 45$, $\Delta S_\text{HL}^{(0)} \approx 5$~cm$^{-1}$/K.}
for Fe\textsuperscript{II}N$_6$-type complexes,
showing excellent correspondence with
the measured relaxation rates
for a variety of SCO and non-SCO Fe(II) compounds
over many orders of magnitude ($10^{-7}$--$10^{2}$~s$^{-1}$).
% LS to MLCT to ... to HS
\subsection{Photoinduced Spin Crossover in Optical Experiments}
\label{sec: SCO-photo-2}
\begin{figure}[ht!]
\centering
\includegraphics[width = \textwidth]{Figures/fig_SCO_MLCT.pdf}
\caption[Calculated ground-state absorption spectrum of BPY.]{
Ground-state absorption spectrum of BPY,
calculated at the CASSCF~level and decomposed by the spin and character
of the final state. Excited states with a dipole moment $< 1.5$~D
and $> 4.0$~D are assigned to MC/LF~(blue) and MLCT~(red) transitions respectively;
others are of mixed character~(gray), involving multiple d--d and MLCT excitations.
Abbreviations: MLCT = metal--ligand charge transfer,
MC = metal-centered, LF = ligand-field.
Adapted with permission from Ref.~\cite{Domingo2014}.
}
\label{fig: SCO-MLCT}
\end{figure}
In the works of this thesis, a specific focus is on
the early-time dynamics of the photoinduced SCO process,
i.e.~the photophysics and structural dynamics
that occur during $\text{LS} \xrightarrow[]{h \nu} \text{HS}$
or, more precisely,
$\mathrm{^1A_{1g}} \xrightarrow[]{h \nu} \mathrm{^1MLCT} \rightarrow \ldots \rightarrow \mathrm{^5T_{2g}}$
where $\ldots$ represents various intermediate states whose identities are still in dispute.
%
Although the $\text{LS} \leftarrow \text{HS}$ relaxation has been
well-understood from the beginning~\cite{Buhks1980, Xie1987, Hauser1991c},
what happens immediately after photoexcitation to the Franck-Condon region
of the $\mathrm{^1MLCT}$ surface (see Fig.~\ref{fig: SCO-MLCT}) and before arrival
to the $\mathrm{^5T_{2g}}$ equilibrium configuration
has proven to be more complex than initially proposed.
Since the discovery of LIESST, a scheme known as the `cascade model'
has been invoked to describe the photophysics driving
the formation of the $\mathrm{^5T_{2g}}$ HS state following photoexcitation of
the $\mathrm{^1A_{1g}}$ LS state~\cite{Hauser1991c}.
%
It follows from Kasha's Rule%
\footnote{See Fn.~\ref{fn: Kasha}.} that
%
\begin{equation}
\begin{aligned}
\tau_\text{IVR} \ll \tau_\text{IC} \lesssim \tau_\text{ISC}
\end{aligned}
\end{equation}
%
whereby the initial excited electronic state should first vibrationally relax,
then radiationlessly decay in sequence to the energetically lowest state of the same spin multiplicity
via internal conversion~(IC), and finally undergo multiple intersystem crossings~(ISC)
to states of different spin multiplicity~\cite{Chergui2015}.
In the simplest cases wherein the initial photoexcited state is low enough in energy that
there is only one intervening state of intermediate spin multiplicity,
this model would predict a minimalist relaxation pathway
as traced by the solid blue arrows in Fig.~\ref{fig: SCO-LIESST}d:
and $\mathrm{^5 T_{2g}} \xrightarrow[]{h \nu} \mathrm{^5 E_{g}} \xrightarrow[]{\text{ISC}} \mathrm{^3 T_{1g}}
\xrightarrow[]{\text{ISC}} \mathrm{^1 A_{1g}}$.
%
Indeed, this double-ISC cascades have proven to be accurate description of rLIESST
in $\mathrm{[Fe^{II}(ptz)_6](BF_4)_2}$ when Marino et~al~\cite{Marino2014}
observed transient ESA with $\tau = 1.7$~ps consistent
with absorption from the $\mathrm{^3 T_{1g}}$ state.
% multiple paths
% lots of energy (2.41~eV or 0.23~MJ/mol)
On the other hand, the mechanism for the forward photoinduced SCO reaction
--- $\mathrm{^1 A_{1g}} \xrightarrow[]{h \nu} \mathrm{^1 MLCT} \rightarrow \ldots
\rightarrow \mathrm{^5T_{2g}}$ --- is not so clear.
%
As seen in Fig.~\ref{fig: SCO-LIESST}d, there are many electronic states
of same and intermediate spin multiplicities
that energetically lie between the $\mathrm{^1MLCT}$ and $\mathrm{^5T_{2g}}$ states,
e.g.~$\mathrm{^1T_{1g}, ^1T_{2g}, ^3T_{1g}, ^3T_{2g}, ^3MLCT}$,
which could be populated.
%
Since $\langle \mathrm{^5T_{2g}} | \hat{H}_\text{SO} | \mathrm{^1A_{1g}} \rangle = 0 $
(see App.~\ref{ap: sco} and Tab.~\ref{tab: sco-so}),
it stands to reason that these intermediate states ought
to be involved and there could be multiple viable pathways, some
of which go back to $\mathrm{^1 A_{1g}}$ and others go forward to $\mathrm{^5T_{2g}}$.
This assertion is evidenced by observation of LIESST following excitation
to $\mathrm{^1T_{1g}, ^1T_{2g}, ^3T_{1g}, ^3T_{2g}}$~\cite{Hauser1991c}.
%
Furthermore, as per the $\tau$-ordering of the cascade model,
each intermediate state ought to be sequentially populated
for long enough time for the molecule to undergo vibrational relaxation to
the state-specific equilibrium nuclear configuration before decaying radiationlessly
to the next state in the cascade,
altogether leading to an overall lengthy relaxation process.
% Problem with cascade model
As it turns out, the cascade model for photoinduced SCO is most likely incorrect,
failing on contact with even the earliest experimental results~\cite{McCusker2003, Juban2006, Zhang2018}.
%
% quantum yield = 1
Before the discovery of LIESST by McGarvey and Lawthers,
Netzel et~al~\cite{Creutz1980, Bergkamp1983} used picosecond TA~spectroscopy
to study the photophysics of $\mathrm{[Fe^{II}(phen)_3]^{2+}}$ and $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$,
reporting that the $\mathrm{^5 T_{2g}}$ state is formed
following $\mathrm{^1 A_{1g}} \xrightarrow[]{h \nu} \mathrm{^1 MLCT}$
and does so with unit efficiency ($\Phi = 1.00 \pm 0.05$).
%
Indeed, such a high quantum yield is hard to reconcile with the imagined cascade through
intermediate singlet and triplet states since they have good coupling
with the $\mathrm{^1 A_{1g}}$ ground state and thus ought to give rise to
multiple relaxation pathways that compete with the SCO process~\cite{SCO-III};
from Ref.~\cite{Hauser1991a}, direct excitation of the $\mathrm{^3 T_{1g}}$ state
yields both $\mathrm{^1 A_{1g}}$ and $\mathrm{^5 T_{2g}}$ states in a ratio of ca.~1:4.
%
% ultrafast
Later, McCusker et~al~\cite{McCusker1992, McCusker1993, Monat2000}
performed single-wavelength sub-ps TA~studies in more Fe(II) complexes
and did not find any transient ESA signal that could be possibly attributed%
\footnote{
The absorption spectrum of the $\mathrm{^5 T_{2g}}$ state
(or any of the other excited states) is not known a~priori for Fe(II) complexes
like $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$ since they do not exhibit thermal SCO.
To assign the ESA in their TA~data, the authors~\cite{Braterman1992, Monat2000, Zalis2011}
used the change in absorbance~$\Delta A(\lambda)$ due to electrochemical oxidation ($\mathrm{[Fe^{II}L_3]^{2+}} \rightarrow \mathrm{[Fe^{III}L_3]^{3+}}$) and
reduction ($\mathrm{[Fe^{II}L_3]^{2+}} \rightarrow \mathrm{[Fe^{II}L_2 (L^-)]^{1+}}$)
as a proxy for that expected of the $\mathrm{^1 MLCT}$ state,
which is purportedly the result of
$\mathrm{(t_{2g})^6 (e_g^*)(L \unslant[-.2]\pi^*)} \rightarrow
\mathrm{(t_{2g})^5 (e_g^*)(L \unslant[-.2]\pi^*)^1}$.}
to any of the intermediate ligand-field states (see Fig.~\ref{fig: SCO-literature-TA}a).
%
Instead, they suggested that the initial $\mathrm{^1 MLCT}$ excited state decays
directly to the $\mathrm{^5 T_{2g}}$ state on an ultrafast time scale~$\tau$
indistinguishable from their IRF time, at first ca.~700~fs~\cite{McCusker1992, McCusker1993}
and then $\lesssim$100~fs~\cite{Monat2000},
followed by $8$-ps kinetics that is attributed to IVR of the final state.
%
\begin{figure}[ht!]
\centering
\includegraphics[width = \textwidth]{Figures/fig_SCO_literature_TA.pdf}
\caption[Time-resolved optical measurements of photoinduced SCO.]{
Time-resolved optical measurements of photoinduced SCO.
(a) Single-wavelength TA scan of $\mathrm{[Fe^{II}(tren(py)_3)](PF_6)_2}$ in acetonitrile
at $\lambda_\text{probe} = 620$~nm following 400-nm excitation;
inset shows the absorption spectra~(blue) of LS~ground state and
the differential absorbance~(red) of the MLCT proxy,
i.e.~the same molecule in its reduced electrochemical state
$\mathrm{[Fe^{II}(L)_2(L^-)]^{2+}}$.
(b) Femtosecond FLUPS spectra~(top)
and broadband TA decay-associated spectra~(bottom)
of aqueous $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$ under 400-nm excitation;
the spectral feature at 466~nm is due to Raman scattering
of the pump light by the vibrational modes of water
($\tilde{\nu} \approx{} 3585$~cm$^{-1}$~\cite{Harris1978}).
(c) Steady-state Raman spectra of
$\mathrm{[Fe^{II}}$(tren(6-\textit{H}-py)$\mathrm{_3)](PF_6)_2}$~(left, blue)
and $\mathrm{[Fe^{II}}$(tren(6-$\mathrm{CH_3}$-py)$\mathrm{_3)](PF_6)_2}$~(right, red)
in acetonitrile, wherein the latter is the HS proxy of the former;
on the right side, the time evolution of the position of the spectral feature near $1650$~cm$^{-1}$
after $560$-nm excitation of the un-methylated complex.
(d) Broadband TA~spectra of aqueous $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$
under 580-nm excitation with UV~(top left) and Vis~(bottom left) probes;
the $311$-nm time trace~(right, black) is shown with its fitted model components~(coloured, solid)
and a slower 120--130-fs time profile as reported by others.
Adapted from Refs.~\cite{Monat2000, Gawelda2007a, Smeigh2008, Aubock2015}
respectively.
}
\label{fig: SCO-literature-TA}
\end{figure}
The absence of ligand-field intermediate states and an intersystem crossing
that is faster than IVR obviate the initially proposed cascade model of photoinduced SCO.
However, a direct $\mathrm{^1 MLCT} \rightarrow{} \mathrm{^5 T_{2g}}$ transition
is unlikely as well.
%
As noted by Refs.~\cite{McCusker2003, SCO-III, Juban2006},
beyond the lack of any spin--orbit coupling between
the $\mathrm{^1 MLCT}$ and $\mathrm{^5 T_{2g}}$ states~(see Tab.~\ref{tab: sco-so}),
the driving force for such a two-electron transition is expected to be
deep in the Marcus inverted region%
\footnote{Canadian chemist Rudolph A. Marcus~(1923--present) developed in 1956
a theory of electron transfer amongst molecules which predicts that
the rate constant~$k$ of such reactions do not increased monotonically with
the total Gibbs free energy~$\Delta{} G^\circ$ but
has a Gaussian dependence~\cite{Nobel1991, Marcus1993}.}
and should not give rise to a rate constant~$k = 1/\tau$ in excess of $10^{12}$~s$^{-1}$
as observed.
To check for short-lived intermediate states that could have been missed by
earlier single-wavelength TA~experiments,
Gawelda et~al~\cite{Gawelda2007a} employed
femtosecond fluorescence up-conversion spectroscopy~(FLUPS)
and broadband TA~spectroscopy (see Fig.~\ref{fig: SCO-literature-TA}b).
%
In their time-resolved emission spectra (Fig.~\ref{fig: SCO-literature-TA}b, top),
a clear signal with $\tau_1 = (20 \pm 5)$~fs is seen in the 600-nm region
where luminescence from the $\mathrm{^1 MLCT}$ state can be expected;
singular value decomposition of the broadband TA~data reveals
a spectral component which decays over $\tau_2 = (116 \pm 10)$~fs and resembles
the static absorption spectrum of the singly-reduced species of $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$,
a known proxy for MLCT states.
%
Earlier studies~\cite{Damrauer1997, Yeh2000, Cannizzo2006}
into the excited-state dynamics of $\mathrm{[Ru^{II}(bpy)_3]^{2+}}$,
a transition-metal complex that is isoelectronic to $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$,
have used the same techniques and
observed similar emission and absorption features that were explained by
a $\lesssim$20-fs decay of the initial $\mathrm{^1 MLCT}$ excited state
and ca.~100-fs population of the $\mathrm{^3 MLCT}$ state.%
\footnote{The evolution into the $\mathrm{^3 MLCT}$ state happens to be much more easily observable
in $\mathrm{[Ru^{II}(bpy)_3]^{2+}}$ because of its strong luminescence
and very long lifetime ($\tau = 630$~ns at $298$~K in water),
properties made possible by the fact that
$\Delta_\text{oct}(\text{Ru}) \approx 2 \Delta_\text{oct}(\text{Fe})$
for a given ligand (see Tab.~\ref{tab: cft-full} in App.~\ref{ap: cft})
and the ligand-field excited states are thus too high in energy to effectively
quench the MLCT states~\cite{Thompson2013, Hauser2017}.}