-
Notifications
You must be signed in to change notification settings - Fork 0
Commit
This commit does not belong to any branch on this repository, and may belong to a fork outside of the repository.
- Loading branch information
Showing
23 changed files
with
19,114 additions
and
0 deletions.
There are no files selected for viewing
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
Original file line number | Diff line number | Diff line change |
---|---|---|
@@ -0,0 +1,64 @@ | ||
|
||
\chapter{Crystal Field Splitting Energy} | ||
\label{ap: cft} | ||
|
||
Table~\ref{tab: cft-full} lists representative values of | ||
the crystal field splitting energy~$\Delta_\text{oct}$ | ||
for octahedral complexes of the form $\mathrm{[ML_6]}^{n+}$ | ||
for monodentate ligands L and $\mathrm{[ML_3]}^{n+}$ for bidentate ones. | ||
The ligands are sorted according to the spectrochemical series, | ||
in increasing order of ligand field strength~\cite{Ryutaro1938a, Ryutaro1938b, FiggisBook}. | ||
% | ||
\begin{table}[ht!] | ||
\centering | ||
{\renewcommand*{\arraystretch}{1.5} | ||
\begin{tabular}{ c c c c c c c c c c } | ||
\toprule | ||
\multirow{2}{*}{ | ||
\begin{minipage}[c]{1.75cm} \centering Electron System \end{minipage} | ||
} & \multirow{2}{*}{Ion} | ||
& \multicolumn{8}{c }{$\Delta_\text{oct}$ ($\times 10^3$~cm$^{-1}$)} \\ | ||
\cline{3-10} | ||
& & Br$^-$ & Cl$^-$ & F$^-$ & ox$^{2-}$ & OH$_2$ & NH$_3$ & en & CN$^-$ \\ | ||
\midrule | ||
d$^1$ & Ti$^{3+}$ & & 13 & 18.9 & & 20.1 & 17 & & 22 \\ | ||
& V$^{4+}$ & & 15 & 20.1 & & & & & \\ | ||
d$^2$ & V$^{3+}$ & & 12 & 16.1 & 18 & 20 & 18 & & 24 \\ | ||
d$^3$ & V$^{2+}$ & & 8 & & & 12.4 & & & \\ | ||
& Cr$^{3+}$ & 13 & 13 & 14.5 & 17.4 & 17 & 21.5 & 22 & 26 \\ | ||
& Mo$^{3+}$ & 14.5 & 19.2 & & & & & & \\ | ||
d$^4$ & Cr$^{2+}$ & & 10 & 14 & & 13 & & 18 & \\ | ||
& Mn$^{3+}$ & & 17.5 & 22 & 20 & 20 & & & 31 \\ | ||
d$^5$ & Mn$^{2+}$ & 7 & 7.5 & 7.8 & & 8.5 & & & 33 \\ | ||
& Fe$^{3+}$ & & & 14 & 14 & 14 & & & 35 \\ | ||
d$^6$ & Fe$^{2+}$ & & & 10 & & 10 & & & 32 \\ | ||
& Co$^{3+}$ & & & 13 & 18 & 20.8 & 22.9 & 23.2 & \\ | ||
& Ru$^{2+}$ & & & & & 19.8 & 28.1 & & \\ | ||
& Rh$^{2+}$ & 19 & 20.4 & & 26 & 27 & 34 & 35 & \\ | ||
& Ir$^{3+}$ & 23 & 25 & & & & 41 & 41 & \\ | ||
& Pt$^{4+}$ & 25 & 29 & 33 & & & & & \\ | ||
d$^7$ & Co$^{2+}$ & 6.5 & 7.65 & 8.3 & 11 & 9.3 & 10.2 & 11 & \\ | ||
d$^8$ & Ni$^{2+}$ & 6.8 & 7.2 & 7.25 & & 8.5 & 11.5 & 11.5 & \\ | ||
d$^9$ & Cu$^{2+}$ & & & & & 12 & 16 & 16 & \\ | ||
\bottomrule | ||
\end{tabular} | ||
} | ||
\caption{Octahedral CFT splitting energy~$\Delta_\text{oct}$ | ||
of some transition-metal complexes of the form [ML\textsubscript{6}]\textsuperscript{n+} | ||
for all monodentate ligands L. Ligands $\mathrm{L = ox^{2-}}$ and en are bidentate | ||
and form complexes of the form $\mathrm{[ML_3]}^{n+}$. | ||
Note that $\mathrm{ox^{2-} = (COO)_2^{2-}}$ is the oxalate dianion | ||
and $\mathrm{en = C_2 H_4 (NH_2)_2}$ is ethylenediamine. | ||
Data herein is taken from Ref.~\cite{FiggisBook}. | ||
} | ||
\label{tab: cft-full} | ||
\end{table} | ||
|
||
Note that values of $\Delta_\text{oct}$ also follow an increasing trend as a function | ||
of the oxidation state and period number of the metal cation, | ||
e.g. $\Delta_\text{oct}(\mathrm{Fe^{3+}}) < \Delta_\mathrm{o}(\mathrm{Fe^{2+}})$ | ||
and $\Delta_\mathrm{o}(\mathrm{Fe^{2+}}) < \Delta_\mathrm{o}(\mathrm{Ru^{2+}})$. | ||
Both effects are consequences of $\Delta_\text{oct} \propto \frac{1}{r^5}$, | ||
where $r$ is the metal--ligand interatomic distance; | ||
a more highly charged metal cation pulls its ligands closer while | ||
a larger one allows them to approach by lessening steric hindrance~\cite{FiggisBook}. |
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
Original file line number | Diff line number | Diff line change |
---|---|---|
@@ -0,0 +1,86 @@ | ||
\chapter{Derivation of the Mott-Bethe Formula} | ||
\label{ap: MottBethe} | ||
|
||
The Mott-Bethe formula is often used to convert in between | ||
the X-ray scattering factor | ||
% | ||
\begin{equation} | ||
f_\text{X}(\boldsymbol{q}) = \int n_\text{e}(\boldsymbol{r}) \mathrm{e}^{ - \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r | ||
\end{equation} | ||
% | ||
and the electron scattering factor | ||
% | ||
\begin{equation} | ||
f_\text{e}(\boldsymbol{q}) = - \frac{m_\text{e}}{2 \unslant[-.2]\pi \hbar^2} \int U(\boldsymbol{r}) \mathrm{e}^{ - \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r | ||
\end{equation} | ||
% | ||
for the purpose of calculations. | ||
However, the source of this handy relationship is rarely stated. | ||
Thus, for the completeness of this thesis, | ||
the derivation for this equation is described below, | ||
as laid out in Ref.~\cite{LandauLifshitzBook}. | ||
|
||
During a scattering event with an atom, X-rays interact mainly with | ||
the electron number density $n_\text{e}(\boldsymbol{r})$ | ||
while electrons are deflected by the electrostatic potential $\phi(\boldsymbol{r})$ | ||
caused by the bound electrons and the nucleus. | ||
% | ||
A starting point to relate these two quantities is the Poisson Equation | ||
and the total atomic charge density~$\rho(\boldsymbol{r})$, | ||
% | ||
\begin{equation} | ||
\nabla^2 \phi(\boldsymbol{r}) = -\frac{\rho(\boldsymbol{r})}{\epsilon_0} | ||
\end{equation} | ||
% | ||
in which their Fourier transforms are inserted, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
\nabla^2 \left( \int \phi_{\boldsymbol{q}} \mathrm{e}^{ \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r} } \mathrm{d}^3 q \right) | ||
& = -\frac{1}{\epsilon_0} \left( \int \rho_{\boldsymbol{q}} \mathrm{e}^{\mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 q \right) \\ | ||
\int \phi_{\boldsymbol{q}} \left( -q^2 \mathrm{e}^{\mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \right) \mathrm{d}^3 q | ||
& = -\frac{1}{\epsilon_0} \int \rho_{\boldsymbol{q}} \mathrm{e}^{\mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 q \\ | ||
\phi_{\boldsymbol{q}} & = \frac{1}{\epsilon_0 q^2} \rho_{\boldsymbol{q}} | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
Therefore, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
\int \phi(\boldsymbol{r}) \mathrm{e}^{ - \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r | ||
& = \frac{1}{\epsilon_0 q^2} \int \rho(\boldsymbol{r}) \mathrm{e}^{ - \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
On the right side, $\rho(\boldsymbol{r})$ can be expanded into a point-like nuclear charge density | ||
$\rho_\text{N}(\boldsymbol{r}) = Z e \delta^3(\boldsymbol{r})$ and an extended electronic charge density $\rho_\text{e}(\boldsymbol{r}) = e n_\text{e}(\boldsymbol{r})$, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
\frac{1}{\epsilon_0 q^2} \int \rho(\boldsymbol{r}) \mathrm{e}^{\mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r | ||
& = \frac{1}{\epsilon_0 q^2} \int \left( Z e \delta^3(\boldsymbol{r}) - e n_\text{e}(\boldsymbol{r}) \right) \mathrm{e}^{\mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r \\ | ||
& = \frac{e}{\epsilon_0 q^2} \left( Z - \int n_\text{e}(\boldsymbol{r}) \mathrm{e}^{\mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r \right) \\ | ||
& = \frac{e}{\epsilon_0 q^2} \left( Z - f_\text{X}(\boldsymbol{q}) \right) | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
where $Z$ is the atomic number of the atom and $n_\text{e}(\boldsymbol{r})$ is the number density of the bound electrons. | ||
% | ||
On the left side, recall that $U(\boldsymbol{r}) = -e \phi(\boldsymbol{r})$, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
\int \phi(\boldsymbol{r}) \mathrm{e}^{ - \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r | ||
& = -\frac{1}{e} \int U(\boldsymbol{r}) \mathrm{e}^{ - \mathrm{i} \boldsymbol{q} \cdot \boldsymbol{r}} \mathrm{d}^3 r \\ | ||
& = \frac{2 \unslant[-.2]\pi \hbar^2}{m_\text{e} e} f_\text{e}(\boldsymbol{q}) | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
By recombining the result from the left and right sides, the Mott-Bethe formula is thus recovered: | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
f_\text{e}(\boldsymbol{q}) = \frac{m_\text{e} e^2}{2 \unslant[-.2]\pi \hbar^2 \epsilon_0 q^2} \left( Z - f_\text{X}(\boldsymbol{q}) \right) | ||
\end{aligned} | ||
\end{equation} |
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
Original file line number | Diff line number | Diff line change |
---|---|---|
@@ -0,0 +1,165 @@ | ||
|
||
\chapter{Theory of Spin Crossover} | ||
\label{ap: sco} | ||
|
||
In Section~\ref{sec: SCO-theory}, the theory of thermal and photoinduced | ||
spin crossover is discussed. | ||
% | ||
In particular, the non-adiabatic multiphonon relaxation model | ||
by Buhks et al~\cite{Buhks1980} is mentioned in the context of | ||
the LS$\leftarrow$HS relaxation following LIESST. | ||
% | ||
Here, this model is described in more detail for reference. | ||
|
||
Following the lead of Hauser in Ref.~\cite{SCO-II}, | ||
consider Fermi's Golden Rule~\cite{Dirac1927, Fermi1950}, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
w_{i \rightarrow f} & = \frac{2 \unslant[-.2]\pi}{\hbar} |V_{i f}|^2 \rho_f | ||
\end{aligned} | ||
\label{eq: fermi-golden} | ||
\end{equation} | ||
% | ||
where $w_{i \rightarrow f}$ is the probability of transition from an initial state~$i$ | ||
(in the $m$-th vibrational level of the HS state) | ||
to a final state~$f$ (in the $m^\prime$-th vibrational level of the LS state), | ||
$|V_{i f}|^2$ is the matrix element of the perturbing potential | ||
that couples states $i$ and $f$, and $\rho_f$ is the density of states at $f$. | ||
|
||
Given that the perturbing potential is | ||
the spin--orbit interaction~$\hat{H}_\text{SO} = \zeta \hat{\boldsymbol{L}} \cdot \hat{\boldsymbol{S}}$, | ||
the coupling matrix element can be expanded as | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
|V_{i f}|^2 & = |\langle \Psi_f | \hat{H}_\text{SO} | \Psi_i \rangle|^2 \\ | ||
& = |\langle \psi_\text{HS} | \hat{H}_\text{SO} | \psi_\text{LS} \rangle|^2 | ||
| \langle \chi_{m^\prime} | \chi_m \rangle |^2 | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
where $| \psi \rangle, | \chi \rangle $ are the electronic and vibrational parts | ||
of the total wavefunction~$| \Psi \rangle$ and, | ||
according to the Condon~approximation, $| \Psi_i \rangle = | \psi_\text{HS} \rangle | \chi_m \rangle$ | ||
and $| \Psi_f \rangle = | \psi_\text{LS} \rangle | \chi_{m^\prime} \rangle$. | ||
% | ||
To evaluate this quantity, consider that | ||
the electronic states are mixed to some degree | ||
due to the presence of the spin--orbit interaction~$\hat{H}_\text{SO}$ | ||
and thus need to be expressed using first-order perturbation theory, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
| \psi_\text{LS} \rangle | ||
& \approx | \psi_\text{LS}^{(0)} \rangle | ||
+ \sum_{j} | ||
\frac{\langle \psi_j^{(0)} | \hat{H}_\text{SO} | \psi_\text{LS}^{(0)} \rangle}{E_\text{LS}^{(0)} - E_j^{(0)}} | \psi_j^{(0)} \rangle \\ | ||
\langle \psi_\text{HS} | | ||
& \approx \langle \psi_\text{HS}^{(0)} | | ||
+ \sum_{j} | ||
\frac{\langle \psi_\text{HS}^{(0)} | \hat{H}_\text{SO} | \psi_j^{(0)} \rangle}{E_\text{HS}^{(0)} - E_j^{(0)}} \langle \psi_j^{(0)} | | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
where the energy denominators are evaluated in | ||
the equilibrium nuclear configuration of the respective states. | ||
Then, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
\langle \psi_\text{HS} | \hat{H}_\text{SO} | \psi_\text{LS} \rangle | ||
& = \langle \psi_\text{HS}^{(0)} | \hat{H}_\text{SO} | \psi_\text{LS}^{(0)} \rangle | ||
+ \sum_{j} \langle \psi_\text{HS}^{(0)} | \hat{H}_\text{SO} | \psi_j^{(0)} \rangle | ||
\langle \psi_j^{(0)} | \hat{H}_\text{SO} | \psi_\text{LS}^{(0)} \rangle \\ | ||
& \quad \left( \frac{1}{E_\text{LS}^{(0)} - E_j^{(0)}} + \frac{1}{E_\text{HS}^{(0)} - E_j^{(0)}} \right) \\ | ||
& = 0 + \langle \mathrm{^5T_{2g}} | \hat{H}_\text{SO} | \mathrm{^3T_{1g}} \rangle | ||
\langle \mathrm{^3T_{1g}} | \hat{H}_\text{SO} | \mathrm{^1A_{1g}} \rangle | ||
\left( \frac{1}{\Delta E_\text{LI}^{(0)}} + \frac{1}{\Delta E_\text{HI}^{(0)}} \right) | ||
\end{aligned} | ||
\end{equation} | ||
% | ||
where $\Delta E_\text{LI}^{(0)}, \Delta E_\text{HI}^{(0)}$ are the energy difference | ||
between the triplet intermediate state $\mathrm{^3T_{1g}}$ and the other states. | ||
From Ref.~\cite{Griffith1964}, $\mathrm{^3T_{1g}}$, | ||
with electronic configuration $\mathrm{(t_{2g})^5 (e_g^*)^1}$, | ||
is the only term which has non-vanishing spin--orbit matrix elements | ||
with both $\mathrm{^1A_{1g}}$ and $\mathrm{^5T_{2g}}$, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
\langle \mathrm{^5T_{2g}} | \hat{H}_\text{SO} | \mathrm{^1A_{1g}} \rangle | ||
& = 0 \\ | ||
\langle \mathrm{^5T_{2g}} | \hat{H}_\text{SO} | \mathrm{^3T_{1g}} \rangle | ||
& = - \sqrt{6} \zeta \\ | ||
\langle \mathrm{^3T_{1g}} | \hat{H}_\text{SO} | \mathrm{^1A_{1g}} \rangle | ||
& = \sqrt{3} \zeta | ||
\end{aligned} | ||
\end{equation} | ||
|
||
Assume that the LS and HS potentials are identical and harmonic of | ||
the same frequency~$\omega$, with the latter displaced energetically by $\Delta E_\text{HL}^{(0)}$ and | ||
configurationally by $\Delta Q_\text{HL} = \sqrt{6} \Delta r_\text{HL}$ | ||
along a single internal vibrational coordinate~$Q$, namely the totally symmetric metal--ligand stretch mode. | ||
Then, energy conservation requires simply $m^\prime = m + n$, | ||
where $n = \frac{\Delta E_\text{HL}^{(0)}}{\hbar \omega}$ is the reduced energy gap between the LS and HS states, | ||
and the density of states~$\rho_f$ becomes $\frac{g_f}{\hbar \omega}$, | ||
where $g_f = 1$ is the degeneracy of the final electronic state. | ||
|
||
To obtain the LS$\leftarrow$HS relaxation rate constant~$k_\text{HL}(T)$, | ||
Eq.~\eqref{eq: fermi-golden} is combined with those above and | ||
ensemble-averaged over all $m$, | ||
% | ||
\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} | ||
\end{equation} | ||
% | ||
where $\bar{F}_n$ is the ensemble average of the Franck-Condon factor~$F_{m n}$, | ||
% | ||
\begin{equation} | ||
\begin{aligned} | ||
F_{m n}(T) & = | \langle \chi_{m + n} | \chi_m \rangle |^2 \\ | ||
\bar{F}_n(T) | ||
& = \frac{\sum \limits_m F_{m n}(T) \mathrm{e}^{-m \hbar \omega / k_\text{B} T}}{\sum \limits_m \mathrm{e}^{-m \hbar \omega / k_\text{B} T}} | ||
\end{aligned} | ||
\end{equation} | ||
|
||
% Sousa, de Graaf calculations | ||
In a similar procedure, Sousa et~al~\cite{Sousa2013} have evaluate | ||
more spin--orbit coupling matrix elements using high-level quantum-chemical methods. | ||
% | ||
Table~\ref{tab: sco-so} shows some of these computational results. | ||
% | ||
\begin{table}[ht!] | ||
\centering | ||
{\renewcommand*{\arraystretch}{1.5} | ||
\begin{tabular}{| c | c | c c c c c c c c |} | ||
\cline{3-10} | ||
\multicolumn{2}{c|}{} & \multicolumn{8}{c |}{$\psi_f$} \\ | ||
\cline{3-10} | ||
\multicolumn{2}{c|}{} & $\mathrm{^1 A_{1g}}$ | ||
& $\mathrm{^1 T_{1g}}$ & $\mathrm{^1 MLCT}$ & $\mathrm{^3 T_{1g}}$ | ||
& $\mathrm{^3 T_{2g}}$ & $\mathrm{^3 MLCT}$ & $\mathrm{^5 T_{2g}}$ & $\mathrm{^5 MLCT}$ \\ | ||
\hline | ||
\multirow{8}{*}{$\psi_i$} & $\mathrm{^1 A_{1g}}$ & & & & 527.7 & 83.7 & 81.6 & 0 & 0 \\ | ||
& $\mathrm{^1 T_{1g}}$ & & & & 75.5 & 131.4 & 164.7 & 0 & 0 \\ | ||
& $\mathrm{^1 MLCT}$ & & & & 96.0 & 214.3 & 199.9 & 0 & 0 \\ | ||
& $\mathrm{^3 T_{1g}}$ & 527.7 & 75.5 & 96.0 & & & & 417.7 & \\ | ||
& $\mathrm{^3 T_{2}}$ & 83.7 & 131.4 & 214.3 & & & & 219.9 & \\ | ||
& $\mathrm{^3 MLCT}$ & 81.6 & 164.7 & 199.9 & & & & 6.2 & 344.3 \\ | ||
& $\mathrm{^5 T_{2g}}$ & 0 & 0 & 0 & 417.7 & 219.9 & 6.2 & & \\ | ||
& $\mathrm{^5 MLCT}$ & 0 & 0 & 0 & & & 344.3 & & \\ | ||
\hline | ||
\end{tabular} | ||
} | ||
\caption{Select spin--orbit coupling matrix elements | ||
$\langle \psi_f | \hat{H}_\text{SO} | \psi_i \rangle$ | ||
of $\mathrm{[Fe^{II}(bpy)_3]^{2+}}$ in the equilibrium nuclear configuration | ||
of the $\mathrm{^1 A_{1g}}$ LS state, calculated at the CASSCF/CASPT2 level~\cite{Sousa2013}. | ||
} | ||
\label{tab: sco-so} | ||
\end{table} |
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
Original file line number | Diff line number | Diff line change |
---|---|---|
@@ -0,0 +1,33 @@ | ||
% \chapter{SVD Analysis of $\mathrm{[Fe^{II}(bpy)_3](PF_6)_2}$ TA Data} | ||
\chapter{SVD Analysis of [Fe\textsuperscript{II}(bpy)\textsubscript{3}](PF\textsubscript{6})\textsubscript{2} TA Data} | ||
\label{ap: sco-bpy} | ||
|
||
% Figure: aqueous SVD analysis | ||
\begin{figure}[hp] | ||
\centering | ||
\includegraphics[width = \textwidth]{Figures/fig_BPY_data_aqueous_svd.pdf} | ||
\caption[SVD analysis of solvated BPY TA data.]{ | ||
SVD analysis of solvated BPY TA data: | ||
(a) UV short-time, (b) UV long-time, (c) Vis short-time, and (d) Vis long-time. | ||
From left to right, the panels show | ||
the principal wavelenght-dependent singular vectors~$u_i(\lambda)$, | ||
the first 100 singular values~$s_i$, | ||
and the principal time-dependent singular vectors~$v_i(t)$. | ||
} | ||
\label{fig: BPY-data-aqueous-svd} | ||
\end{figure} | ||
|
||
% Figure: aqueous SVD analysis | ||
\begin{figure}[p] | ||
\centering | ||
\includegraphics[width = \textwidth]{Figures/fig_BPY_data_crystal_svd.pdf} | ||
\caption[SVD analysis of single-crystal BPY TA data.]{ | ||
SVD analysis of single-crystal BPY TA data: | ||
(a) UV short-time, (b) UV long-time, (c) Vis short-time, and (d) Vis long-time. | ||
From left to right, the panels show | ||
the principal wavelenght-dependent singular vectors~$u_i(\lambda)$, | ||
the first 100 singular values~$s_i$, | ||
and the principal time-dependent singular vectors~$v_i(t)$. | ||
} | ||
\label{fig: BPY-data-crystal-svd} | ||
\end{figure} |
Oops, something went wrong.