electronic supporting information ultrafast observation of

36
S1 Electronic Supporting Information Ultrafast Observation of a Photoredox Reaction Mechanism: Photo-initiation in Organocatalyzed Atom-Transfer Radical Polymerization Daisuke Koyama, Harvey J. A. Dale and Andrew J. Orr-Ewing* School of Chemistry, University of Bristol, Cantock’s Close, Bristol, BS8 1TS, United Kingdom * Author for correspondence: [email protected] Contents Additional Analysis PCF in DCM: Assigning 1 PCF*(S 1 ) vs 3 PCF*(T 1 ) PCH and MBP in DMF: Complexation of 2 PC .+ and Br - Experimental Ultrafast laser system Transient vibrational absorption spectroscopy Transient electronic absorption spectroscopy Photocatalyst synthesis Steady-state spectroscopy Computational Methodology Ground-state calculations Excited-state calculations Steady-State Spectroscopic Characterizations FTIR spectra: PCF Electronic spectra: PCF

Upload: others

Post on 27-Jan-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Electronic Supporting Information Ultrafast Observation of

S1

Electronic Supporting Information

Ultrafast Observation of a Photoredox Reaction Mechanism: Photo-initiation

in Organocatalyzed Atom-Transfer Radical Polymerization

Daisuke Koyama, Harvey J. A. Dale and Andrew J. Orr-Ewing*

School of Chemistry, University of Bristol, Cantock’s Close, Bristol, BS8 1TS, United Kingdom

* Author for correspondence: [email protected]

Contents

Additional Analysis

PCF in DCM: Assigning 1PCF*(S1) vs

3PCF*(T1)

PCH and MBP in DMF: Complexation of 2PC

.+ and Br

-

Experimental

Ultrafast laser system

Transient vibrational absorption spectroscopy

Transient electronic absorption spectroscopy

Photocatalyst synthesis

Steady-state spectroscopy

Computational Methodology

Ground-state calculations

Excited-state calculations

Steady-State Spectroscopic Characterizations

FTIR spectra: PCF

Electronic spectra: PCF

Page 2: Electronic Supporting Information Ultrafast Observation of

S2

Spectroscopic characterization of 2PCF

.+(D1)

Spectroscopic characterization of 2PCH

.+(D1)

Computational Figures

Natural transition orbitals: PCF

Initial active space orbitals: PCF

Natural transition orbitals: PCH

Vertical transition energies: PCF

Vertical transition energies: PCH

Scaled harmonic frequencies

Reactivity of the debrominated radical 2MP

.(D1)

Transient Absorption Spectra

Example spectral decompositions

TEAS basis functions

Solvent effects: TEAS kinetic traces

Solvent effects: Semi-logarithmic plots of TEAS kinetic traces

Semi-logarithmic plots for TVAS measurements in DCM solutions

PCH and MBP in DCM: TVA spectra

PCF and MBP in DCM: TEAS concentration dependence

PCH: TEA spectra

Pure DCM: TEA spectra

Emission spectra: PCF with/without triplet quenchers

Triplet sensitization: PCF in DMF with benzophenone

Page 3: Electronic Supporting Information Ultrafast Observation of

S3

Additional Analysis

PCF in DCM: Assigning 1PCF*(S1) vs

3PCF*(T1)

Transient electronic absorption spectra of PCF in DCM following pulsed photoexcitation at 370 nm are

shown in Figure S1. Mechanistic insight into the photophysical relaxation pathway may be garnered from

the early-time kinetic regime (0 – 5 ps). Global fitting of time-dependent integrated band intensities to a

simple A →B (τ1) mechanism, convoluted with an instrument response function of 120 fs, affords a time

constant τ1 = 230 ± 30 fs.

The transient species B is fully formed within 2 ps and metastable on the timescale 2 – 100 ps, but decays

by 30% over 100 – 1200 ps (Figure S2). The time constant for the decay is beyond the range of our

experiment but is estimated to be 3 ± 1 ns.

Figure S1: Ultrafast photophysical relaxation of PCF in DCM. Transient electronic absorption spectra of PCF

(2.1 mmol dm-3

) in DCM are shown for the first 5 ps after photoexcitation at 370 nm, alongside kinetic traces for

1PCF*(Sn) (○) and

1PCF*(S1)(○). The lines show the result of global fitting to a mechanism

1PCF*(Sn) →

1PCF*(S1), convoluted with an instrument response function of 120 ps. Spectral decompositions for example time

slices are shown in Figure S11. Transient signal from pure DCM is negligible under the same conditions and

over the same probe region (see Figure S21).

We assign A as the nascent excited electronic state of the photocatalyst 1PCF*(S4), or transiently

populated S3 or S2 states, and the long-lived state B as the first excited singlet state 1PCF*(S1). The only

alternative electronic state that might plausibly have a lifetime consistent with B is 3PCF*(T1), but such an

assignment was discounted for the following reasons: (i) if intersystem crossing (ISC) from the singlet

manifold to 3PCF*(T1) were to occur with a time constant of hundreds of femtoseconds, the steady-state

emission would be dominated by phosphorescence, rather than fluorescence; (ii) there is no compelling

1PCF*(S1)

1PCF*(Sn (n = 2 - 4))

400 450 500 550 6000

5

10

15

Time / ps

0.0

0.3

0.7

1.5

2.5

4.0

Absorbance / m

OD

Wavelength / nm

-1 0 1 2 3 4 5

0.0

0.2

0.4

0.6

0.8

1.0

Integrated intensity

Time / ps

Page 4: Electronic Supporting Information Ultrafast Observation of

S4

theoretical explanation for such a fast ISC in this structurally rigid molecule, composed exclusively of

first-row main group elements; and (iii) the decay time constant of B (3 ±1 ns) is far shorter than is often

observed for phosphorescence or ISC from the first excited triplet state to the ground state of organic

molecules. Assigning the metastable state B to 3PCF*(T1) would therefore require two abnormally fast

ISCs – one from a high-lying singlet state into the triplet manifold, and a second from 3PCF*(T1) to the

ground state 1PCF(S0). Moreover, none of the triplet quenchers Q considered in the steady-state

experiments (2,5-dimethyl-hexa-2,4-diene, cyclohexa-1,3-diene, styrene) had any effect on the decay

kinetics of B on our timescale, even when present in vast excess ([Q] ~1 mol dm-3

).

Although global fitting to a sequential mechanism A →B → C proved intractable, the presence of at least

two independent absorption features between 400 – 500 nm is supported by the asymmetric spectral

profile in this region, which undergoes a concerted narrowing and red-shift over the first few picoseconds.

This, and the shifting isosbestic point over 500 – 510 nm, suggest the involvement of an intermediate

singlet state and therefore a stepwise, rather than concerted, descent from 1PCF*(S4) to

1PCF*(S1) via

sequential internal conversions.

The transient vibrational absorption spectra of PCF over 1280 – 1380 cm-1

(Figure S2) depict vibrational

signatures of 1PCF*(S1) and the ground-state

1PCF(S0), corresponding to fundamental transitions in ring-

breathing modes. The early-time spectra are characteristic of vibrational relaxation, and the temporal

evolution of the integrated band intensities affords a time constant of 15 ± 2 ps.

Page 5: Electronic Supporting Information Ultrafast Observation of

S5

Figure S2: Photophysical dynamics of PCF in DCM following pulsed photoexcitation at 370 nm. (a) TVA

spectra showing vibrational cooling of the 1PCF*(S1) state, decay of

1PC*(S1) and recovery of the ground-state

1PCF(S0). Inset: time-dependence of the

1PCF*(S1) band position, affording a time constant of (15 ± 2 ps) for

vibrational relaxation; (b) kinetic traces for depletion of 1PCF*(S1) (black) and recoveries for the

1PCF(S0) bands

at 1324 cm-1

(blue) and 1345 cm-1

(red); (c) TEA spectra of the same system, showing the decay of the 1PCF*(S1)

state over the period 10 – 1200 ps; (d) kinetic trace of 1PCF*(S1), as obtained from (c) via the time-dependent

integration of the fitted band. Crucially, the total decay of the 1PCF*(S1) state (30 % over 30 – 1200 ps) is

matched by the fractional recovery of the ground-state; this correspondence suggests that the ISC quantum yield

is imperceptibly low.

The depletion of A and growth of B occur with a time constant (230 ± 30 fs) an order of magnitude

smaller than that found for the vibrational relaxation (15 ± 2 ps) of the 1PCF*(S1) state, suggesting that

the red-shift of the spectra over 400 – 500 nm cannot correspond to vibrationally hot 1PCF*(S1). Solvent

reorganization following a direct internal conversion from 1PCF*(S4) to

1PCF*(S1) might lead to early-

time shifts in the excited state absorptions, but the anticipated stabilization of the 1PCF*(S1) state due to

1300 1320 1340 1360 1380 1400

-20

-10

0

10

20

30Absorbance / m

OD

Wavenumber / cm-1

Time / ps

1

15

50

200

500

(c) TEAS (d)

1PCF*(S1)

1PCF(S0): 1345 cm-1

1PCF(S0): 1324 cm-1

1PCF*(S1)

1PCF*(S1)

(a) TVAS (b)

PCF(S0)

1PCF*(S1)

0 20 40 60 80 1001304

1306

1308

1310

Band position / cm

-1

Time / ps

200 400 600 800 1000 1200

-0.6

-0.4

-0.2

0.6

0.8

1.0

Integrated signal

Time / ps

400 450 500 550 6000

5

10

15

Absorbance / m

OD

Wavelength / nm

Time / ps

10

100

500

800

1200

200 400 600 800 1000 12000.0

0.2

0.4

0.6

0.8

1.0

Integrated signal

Time / ps

Page 6: Electronic Supporting Information Ultrafast Observation of

S6

its CT character should cause a progressive blue-shift, rather than the observed red-shift, of the transient

electronic spectra over the first few ps.

Critically, the total decay amplitude of the excited singlet state over 100 – 1200 ps corresponds closely to

the fractional recovery of the ground-state bleach, suggesting that fluorescence and internal conversion –

but not intersystem crossing – are efficient relaxation pathways for 1PCF*(S1). The formation of

3PCF*(T1) would also afford a distinct transient absorption growing over time, but no such band was

observed in any region. This evidence suggests that the ISC quantum yield is imperceptibly low, and that

a substantial majority of 1PCF*(S1) returns to the ground-state within several nanoseconds, in agreement

with the assignment of 1PCF*(S1) from the TEA spectra above (Figure 1).

PCH and MBP in DMF: Complexation of 2PC

.+ and Br

-

In DMF solution there is a component in the transient electronic spectrum of PCH + MBP centered on

470 nm (Figure 4), which continues to grow significantly after 1PCH*(S1) is depleted. As

1PCH*(S1) is

almost fully depleted within 100 ps this feature cannot correspond to intersystem crossing to 1PCH*(T1),

nor can it correspond to a PET from 3PCH*(T1), as the population of the

2MP

.(D1) radical reaches a

plateau after 100 ps. We attribute this subtle change in the absorption signature of the radical cation

2PCH

.+(D1) to complexation with bromide to form the radical ion pair

2PCH

.+Br

-.

Page 7: Electronic Supporting Information Ultrafast Observation of

S7

Experimental

Ultrafast Laser System

Details of the ultrafast laser system and experimental procedure are summarised in the Methods section of

the main paper and elsewhere.1

Transient Vibrational Absorption Spectroscopy (TVAS)

For all TVAS experiments the carrier wavelength of the UV pump pulse was set to 370 nm and the IR

probe pulses were characterised by a bandwidth of approximately 250 cm-1

. The UV pump pulses emitted

from the OPA were attenuated by cross-polarization to afford pulses of energy 600 nJ at the sample. The

UV pump and IR probe pulses were configured to intersect each other at the sample at a small angle of 5°,

and precisely controlled time delays τ between the pump and probe pulses were attained with an

aluminium retro-reflector mounted on an adjustable mechanical delay stage of length 220 mm (Thorlabs

DD5220/M). This corresponds to a maximum temporal delay of 1.3 ns. Difference absorption spectra

∆A(ν, τ) were obtained by placing a revolving chopper wheel along the UV pump beamline, with an

angular frequency of 500 Hz and a phase adjusted to block alternate pump pulses. Prior to interacting with

the sample, the IR probe beam was split into two beams of identical energy: (i) one beam was transmitted

through the sample and directed onto a Mercury Telluride Cadmium array (Infrared Associates Inc.,

MCT-10-128) mounted within an infrared spectrometer (Horiba Scientific, iHR320, 2 cm-1

pixel-1

); and

(ii) the second beam was directed straight into an identical detector and spectrometer, without passing

through the sample. Comparison of the two recorded signals accounted for fluctuations between

successive pulses, thereby reducing artificial differences between individual transient spectra recorded for

any given time delay. Systematic drifts in the signal intensity over the timescale of the experiment can

lead to the emergence of misleading kinetics; to avoid this possibility, transient spectra were recorded in a

randomized – as opposed to ascending or descending – sequence of time delays. To minimise absorption

of the infrared probe by ambient water vapor, the beamline was partially enclosed and purged with dry

nitrogen.

Transient Electronic Absorption Spectroscopy (TEAS)

For TEAS a broadband white light supercontinuum was used to probe the wavelength region 330–630

nm, following photoexcitation with a pump pulse of carrier wavelength 370 nm. The supercontinuum

Page 8: Electronic Supporting Information Ultrafast Observation of

S8

probe pulses were generated by focussing the parent 100 �J pulse beam onto a CaF2 window of 2mm

thickness with a CaF2 lens of focal length 200 mm; to prevent thermal degradation of the window a cross-

polariser and series of neutral density filters were used to attenuate the parent beam intensity and the

window was continuously rastered in the plane perpendicular the incident beam direction. The

polarization of the supercontinuum probe was shifted relative to the UV pump pulse by the magic angle

(54.7°), and the two beams were configured to intersect at the sample at an angle of 5°. Difference

absorption spectra were obtained by placing a revolving chopper wheel along the UV pump beamline,

with an angular frequency of 500 Hz and a phase adjusted to block alternate pump pulses. The transmitted

probe pulses were collected by an optical fibre coupled to a CCD spectrometer (Avantes, AvaSpec-

DUAL), affording a spectral resolution of 0.6 nm pixel-1

.

Photocatalyst Synthesis2

5,10-dihydrophenazine 1

A conical flask was charged with phenazine (2.0 g, 11 mmol) in EtOH (50 mL) and heated to reflux. With

stirring, a solution of Na2S2O4 (23.3 g, 134 mmol) in 200 mL of distilled water was then added over a

period of 5 min, leading to the formation of a pale green precipitate after 10 min. The crude solid was

isolated via vacuum filtration, washed thoroughly with distilled water and dried in vacuo to afford 5,10-

dihydrophenazine (1) as a pale green powder (1.75 g, 9.6 mmol, 87 %). The photocatalysts PCF (2) and

PCH (3) were synthesised immediately after exposing 1 to the atmosphere.

5,10-di(4-trifluoromethylphenyl)-5,10-dihydrophenazine (PCF) 2

The photocatalyst PCF (2) was synthesised via a Buchwald-Hartwig coupling reaction. An oven-dried

round-bottom flask was charged with a magnetic stirrer bar, 1 (0.50 g, 2.7 mmol), NaOtBu (1.0 g, 10.4

mmol), RuPhos (50 mg, 0.11 mmol), RuPhos precatalyst (90mg, 0.11 mmol), para-

bromobenzotrifluoride (2.5 g, 11 mmol) and 1,4-dioxane (6 ml) and heated to 100 °C under an

atmosphere of nitrogen. After 10 hr the crude yellow solid was isolated via vacuum filtration, dissolved in

Na2S2O4

EtOH / H2O

1

Page 9: Electronic Supporting Information Ultrafast Observation of

S9

a minimal quantity of DCM and then crystallised with the addition of n-hexane to afford 2 as a yellow

solid (0.96 g, 2.0 mmol, 76 %). Further crystallization from toluene solution gave yellow crystals. The

crystal was stored below 5°C. Characterization was confirmed by comparison the previously reported

values. δ1H / ppm (toluene-d8, 400 MHz) = 7.37 (d, J = 8.0 Hz, 4H), 6.97 (unable to resolve from the

toluene-d8 peak), 6.31 (m, 4H), 5.65 (m 4H). HRMS (ESI) M+ m/z = 470.1218 (Calc.), 470.1220 (Meas.).

UV-Vis �max (DCM) = 370 nm.

5,10-diphenyl-5,10-dihydrophenazine (PCH) 3

The photocatalyst PCH (3) was synthesised in a similar manner to the synthesis of 2. An oven-dried

round-bottom flask was charged with a magnetic stirrer bar, 1 (0.50 g, 2.7 mmol), NaOtBu (1.0 g, 10.4

mmol), RuPhos (50 mg, 0.11 mmol), RuPhos precatalyst (90mg, 0.11 mmol), para-bromobenzene (1.7 g,

11 mmol) and 1,4-dioxane (6 mL). The solution was heated to 100 °C under an atmosphere of nitrogen

and stirred for 10 hr. After cooling to room temperature, 100 mL of DCM was added to the reaction

solution, and the solution washed with 100 mL of distilled water three times. The organic layer was

collected and concentrated to ~5 mL in vacuo, causing precipitation of a yellow powder that was then

collected via vacuum filtration. This purification procedure was repeated once more, giving 3 as yellow

powder (0.75 g, 2.2 mmol, 83%). Recrystallization from toluene / methanol solution gave yellow needles,

which were stored below 5°C. Characterization was confirmed by comparison to previously reported

values. δ1H / ppm (toluene-d8, 400 MHz) = 7.20 (t, J = 7.6 Hz, 4H), 7.13 (unable to resolve from toluene-

d8 peak), 7.07 (m, 2H), 6.24 (m, 4H), 5.75 (m, 4H). HRMS (ESI) M+ m/z = 334.1470 (Calc.), 334.1474

(Meas.).

+

RuPhosRuPhos Precat.

NatOBu

1,4-dioxane

1

2

Page 10: Electronic Supporting Information Ultrafast Observation of

S10

Steady-State Spectroscopy

All steady-state UV-Vis absorption spectra reported herein were obtained with a Thermo Scientific

Genesys 10S spectrophotometer. Emission spectra were measured with a Varian Cary Eclipse

spectrophotometer using 10 nm excitation and 20 nm emission slits. Solutions were prepared to a

concentration such that they exhibited a maximum absorbance of 0.05 < A(�max) < 0.1 at the excitation

wavelength �max. Both electronic absorption spectra and emission spectra were obtained using glass

cuvettes of path length 1 cm. FTIR spectra were recorded on a Perkin Elmer Spectrum-Two spectrometer,

with solutions contained in a stainless steel Harrick cell with CaF2 windows and sample path length of 50

�m.

Computational Methodology

Ground-State Calculations

Restricted Kohn-Sham density functional theory (RKS-DFT) was employed to investigate the geometries

and harmonic vibrational frequencies of pertinent closed-shell species in their ground electronic state (S0).

Geometries and frequencies of open shell species in their ground doublet state (D1) were obtained within

the unrestricted formalism (UKS-DFT). All KS-DFT calculations were run with the Gaussian09 suite of

programs and a fine numerical integration grid at T = 298.15 K.3

Ground-state geometry optimizations and harmonic vibrational frequency calculations were conducted

with the pure hybrid-GGA functional PBE0 and a polarised and augmented double-zeta basis set 6-

31+G(d) (6d,7f).4 All optimizations were initiated from highly distorted structures devoid of any

symmetry elements. Equilibrium solvation effects were included by implicitly simulating the solvent as a

continuous polarisable medium with the total solute density model (SMD) of Truhlar et al.,5-7

and in the

case of the PBE0 functional long-range London dispersion interactions were simulated with Grimme’s D3

RuPhosRuPhos Precat.

NatOBu

1,4-dioxane+

1

3

Page 11: Electronic Supporting Information Ultrafast Observation of

S11

dispersion correction, supplemented by the Becke-Johnson damping function (GD3BJ).8 The eigenvalues

of the Hessian matrix were checked for each species to confirm its presence at a genuine minimum or

saddle-point (TS) on the potential energy surface. Linear scaling factors for harmonic vibrational

frequencies and associated uncertainties were obtained by comparison to experimental steady-state FTIR

spectra. Predicted vibrational frequencies of the transient species 3PCF*(T1),

2PCF

.+(D1),

2PCF

.-(D1),

2MP

.(D1),

2PCF

.+Br

- and {

1PCF(S0), Br2} were obtained by computing their harmonic frequencies and

adjusting with the scaling factor obtained from this comparison. To enable direct comparison with excited

states (below), molecular electrostatic potentials (MEPs) were computed at the CAM-B3LYP/6-

31+G(d)/SCRF=(SMD, DCM) level of theory.9 Accurate ground-state potential energies were obtained

with unrestricted KS-DFT, unless explicitly stated otherwise, with single-point calculations at the

B2PLYPD3/6-311++G(2d,p)/SCRF=(SMD,DCM)//PBE0-GD3BJ/6-31+G(d)/SCRF=(SMD,DCM) level

of theory.9 The corresponding Gibbs free energies were computed by including the thermal correction

from the optimization level of theory.

The Gibbs free energy change for the PET ∆PETG, C – Br bond dissociation energy of MBP EBD and

Gibbs free activation barrier ∆PETG‡ were computed for each photocatalyst according to the equations

below. All ground-state free energies were computed at the B2PLYPD3/6-311++G(2d,p)/SCRF=(SMD,

DCM)//PBE0-GD3BJ/6-31+G(d)/SCRF=(SMD, DCM) level of theory at 298 K. The energy of the

excited singlet state [1PCF*(S1) or

1PCH*(S1)] in each case was estimated from the wavelength of

maximum emission. This method neglects the Coulombic term present in the Weller equation and the

solvent reorganization energy in the Marcus equation, but these are reasonable first-approximations for a

dissociative ET in which the excited state energy and bond dissociation energy are the dominant terms.

Δ�� ≈ ����.�� + ����.� + ������� − ����� + ������ + ��,�� ����MBP� ≈ ����.� + ����.�� − ������� = 210kJmol*+

Δ��‡ ≈���4 .1 + Δ����� /0

KS-DFT computations predict a C – Br bond dissociation energy of EBD[MBP] = 210 kJ mol-1

for the

initiator. The intrinsic rate coefficient for the electron transfer may be calculated according to the equation

below, where κel is the electronic transmission coefficient, νn is the nuclear vibration frequency related to

the electron transfer and κelνn ~ 1012

dm3 mol

-1 s

-1.10

1234 = 56789 exp =−Δ��‡

>? @

Page 12: Electronic Supporting Information Ultrafast Observation of

S12

Photocatalyst AB,B/

(kJ mol-1

) {nm}

CDEFG / (kJ mol

-1)

CDEFG‡ / (kJ mol

-1)

HIAJ / (dm

3 mol

-1 s

-1)

PCF 213 {562} -95 16 1.6 × 109

PCH 249 {480} -145 5 1.3 × 1011

Although this is an approximate computation of rate coefficients, agreement with experiment is good.

Some degree of fortuitous error cancellation may play a role, but it is notable that the calculation captures

the essential conclusions of our experimental work: (i) PET from 1PCH*(S1) to MBP will be limited by

diffusion; and (ii) PET from 1PCF*(S1) is slower than from

1PCH*(S1), and activation controlled, with a

bimolecular rate coefficient of the order 109 dm

3 mol

-1 s

-1.

Excited-State Calculations

Vertical transition energies, oscillator strengths and excitation amplitudes were computed with time-

dependent density functional theory (TD-DFT) and the second-order perturbation-corrected configuration

interaction singles method (CIS(D)) in Gaussian09.11-13

For TD-DFT calculations a selection of global

hybrid-GGA and range-separated exchange-correlation functionals (EX-C) – including CAM-B3LYP,

wB97XD, LC-wPBE, PBE0, M06-2X, B3LYP and B98 – were used in conjunction with the 6-

311++G(2d,p) (5d, 7f) basis set and non-equilibrium SMD solvation to characterise the excited electronic

states of the photocatalyst in the Franck-Condon region. The Coulomb-attenuated variant of the B3LYP

functional (CAM-B3LYP) has been reported to provide satisfactory description of delocalised π-π*

excited states in organic dyes and of charge-transfer excitations generally,14-16

and this was corroborated

herein by comparison to the steady-state UV-vis spectrum of the photocatalyst. More expansive basis sets,

ranging up to the minimally-augmented triple-zeta correlation-consistent basis set from Dunning and co-

workers (spaug-cc-PVTZ), had a negligible effect on computed vertical transition energies and oscillator

strengths. Natural transition orbitals (NTOs) and simulated UV-vis spectra were thus computed at the

CAM-B3LYP/6-311++g(2d,p)/SCRF=(SMD,DCM,NonEq))//PBE0+GD3BJ/6-

31+G(d)/SCRF=(SMD,DCM) level of theory.

State-averaged complete active space self-consistent field (SA-CASSCF) calculations were pursued to

accurately characterise the vertical electronic transitions of the photocatalyst from an ab initio

perspective. All SA-CASSCF calculations were conducted in vacuo with the 6-311G(d,p) basis set using

the Molpro package (v2010.1).17-19

The input geometries was obtained by imposing C2h and D2h

Page 13: Electronic Supporting Information Ultrafast Observation of

S13

symmetries on the ground-state photocatalysts PCF and PCH, respectively, and reoptimising in

Gaussian09 at the PBE0+GD3BJ/6-31+G(d)/SCRF=(SMD,DCM) level of theory; with the exception of

minor rotations in the trifluoromethyl substituents, the resulting structures were essentially isoenergetic

relative to the fully relaxed species. As such, the trifluoromethyl substituents were treated as free rotors.

Initial canonical molecular orbitals were obtained from a restricted Hartree-Fock single-point calculation

with the 6-311G(d,p) basis set, from which a benchmark active space of 14 electrons in 14 orbitals was

constructed (14,14); the active orbitals comprised the seven highest occupied π orbitals and seven lowest

unoccupied π* orbitals. A second, more compact active space (8,8), including only those occupied π

orbitals localised exclusively on the central phenazine moiety and virtual π* orbitals confined primarily to

the trifluoromethylphenyl substituents, was also examined, alongside an intermediate-sized active space

(10,10). Unless explicitly stated otherwise, the optimised SA-CASSCF orbitals were obtained by

minimising the average energy of: (i) the ground electronic state (Ag) and the first excited state of each

symmetry species (Ag, Au, Bu, Bg) for PCF (C2h); and (ii) the ground electronic state (Ag) and the first four

lowest singlet states (B1g, B2u, Au, B3g) for PCH (D2h). In both cases, each of the five states was afforded

an equal weighting in state-averaging.

Accurate vertical transition energies were obtained using internally-contracted second-order

multireference perturbation theory (CASPT2) and adjusted by -0.1 eV to account for basis-set

incompleteness, in accordance with the benchmarking work of Thiel and co-workers.20

The SA-CASSCF

result was used as a reference wavefunction for the CASPT2 calculation and a small imaginary level shift

of 0.1 Eh was employed to circumvent intruder state problems. Transition dipole moments were obtained

from the SA-CASSCF wavefunctions.

Optimised geometries and molecular electrostatic potentials (MEPs) of excited singlet states were

obtained with TD-DFT at the CAM-B3LYP/6-31+G(d)/SCRF=(SMD,DCM) level of theory.21

Page 14: Electronic Supporting Information Ultrafast Observation of

S14

Steady-State Spectroscopy

FTIR spectra: PCF

Figure S3: Assorted FTIR spectra for PCF with solvent absorption subtracted: (a) MBP (0.18 mol dm-3

) in

toluene-d8; (b) PCF (5.3 mmol dm-3

) in toluene-d8; (c) PCF (5.3 mmol dm-3

) and MBP (0.18 mol dm-3

) in toluene-

d8; (d) difference spectrum (c) – (a). The absence of additional vibrational absorptions for the PCF + MBP system

– and therefore the equivalence of (b) and (d) – discounts the formation of a donor-acceptor complex in the

ground-state.

0.0

0.1

0.2

Absorbance

0.0

0.1

0.2

Absorbance

0.00

0.05

0.10

1400 1450 1500 1550 1600 1650 1700

0.00

0.05

0.10

Wavenumber / cm-1

(a) MBP

(b) PCF

(c) MBP + PC

(d) spectrum (c) – (a)

Absorbance

Page 15: Electronic Supporting Information Ultrafast Observation of

S15

Electronic Spectra: PCF

Figure S4: Assorted electronic spectra for PCF. (a) UV-vis absorption spectra for PCF in acetonitrile (∎),

toluene (∎), ethanol (∎), DCM (∎) and DMF (∎); (b) emission spectra for PCF in acetonitrile (∎), toluene (∎),

ethanol (∎), DCM (∎) and DMF (∎) following excitation at 370 nm; (c) excitation spectrum of PCF in DCM for

emission at 570 nm.

300 350 400 450 500

0.0

0.5

1.0

Norm

alized absorbance

Wavelength / nm

DCM

EtOH

MeCN

toluene

DMF

450 500 550 600 650 700

0.0

0.5

1.0

Norm

alized intensity

Wavelength / nm

DCM

EtOH

MeCN

toluene

DMF

300 350 400 450 500

0.0

0.5

1.0

Norm

alized intensity

Wavelength / nm

(b) Emission

(c) Excitation

(a) UV-vis

Page 16: Electronic Supporting Information Ultrafast Observation of

S16

Spectroscopic Characterization of 2PCF

.+(D1)

Figure S5: Steady-state spectroscopic characterization of the radical cation 2PCF

.+(D1), following thermally-

mediated oxidation of PCF with FeCl3. (a) UV-vis absorption spectra of PCF in DCM following the addition of n

equivalents of FeCl3; (b) FTIR spectra of PCF in DCM following the addition of n equivalents of FeCl3. The

emergence of a vibrational absorption at 1553 cm-1

confirms the assignment of 2PCF

.+(D1) in the transient spectra

presented in Figure 1(c).

Spectroscopic Characterization of 2PCH

.+(D1)

Figure S6: Steady-state spectroscopic characterization of the radical cation 2PCH

.+(D1) following thermally-

mediated oxidation of PCH with FeCl3. (a) UV-vis absorption spectra of PCH in DCM following the addition of

n equivalents of FeCl3; (b) FTIR spectra of PCH in DCM following the addition of n equivalents of FeCl3.

1480 1500 1520 1540 1560 1580

0.00

0.05

0.10

0.15

Wavenumber / cm-1

0 eq

0.5 eq

1.0 eq

2.0 eq

3.0 eq

(a) UV-vis (b) FT-IR

2PCF・+(D1)

2PCF・+(D1)

1PCF(S0)

400 450 500 550 600 650 700

0.0

0.2

0.4

0.6

0.8

Absorbance

Wavelength / nm

0 eq

0.5 eq

1.0 eq

1.5 eq

2.0 eq

2.5 eq

1540 1550 1560

1480 1500 1520 1540 1560 1580

0.00

0.05

0.10

0.15

Absorbance

Wavenumber / cm-1

0 eq

0.5 eq

1.0 eq

2.0 eq

3.0 eq

(a) UV-vis (b) FT-IR

2PCH・+(D1)

2PCH・+(D1)

1PCH(S0)

400 450 500 550 600 650 700

0.0

0.2

0.4

0.6

0.8

1.0

Absorbance

Wavelength / nm

0 eq

0.5 eq

1.0 eq

1.5 eq

2.0 eq

2.5 eq

1540 1550 1560

Page 17: Electronic Supporting Information Ultrafast Observation of

S17

Computational Figures

Natural Transition Orbitals: PCF

CAM-B3LYP/6-311++G(2d,p)/SCRF=(SMD,DCM)//PBE0-GD3BJ/6-31+G(d)/ SCRF=(SMD,DCM)

Figure S7: Natural transition orbitals for vertical electronic transitions of PCF to the first four excited singlet

states, computed at the CAM-B3LYP/6-311++G(2d,p)/SMD(DCM)//PBE0-GD3BJ/6-31+G(d)/SMD(DCM) level

of theory. The symmetry of PCF was constrained to the C2h point group.

Page 18: Electronic Supporting Information Ultrafast Observation of

S18

Initial Active Space Orbitals (SA-CASSCF): PCF

HF/6-311G(d,p)

Figure S8: Canonical molecular orbitals for PCF, computed at the HF/6-311G(d,p)//PBE0-GD3BJ/6-

31+G(d)/SMD(DCM) level of theory, with the system constrained to the C2h point group. These orbitals were

used for subsequent SA-CASSCF/CASPT2 computations, as outlined in the computational methodology section.

The PCH canonical MOs, computed at the same level of theory under D2h symmetry, were very similar to those

for PCF, with minor re-orderings.

Page 19: Electronic Supporting Information Ultrafast Observation of

S19

Natural Transition Orbitals: PCH

CAM-B3LYP/6-311++G(2d,p)/SCRF=(SMD,DCM)//CAM-B3LYP/6-31+G(d)/ SCRF=(SMD,DCM)

Figure S9: Natural transition orbitals for vertical electronic transitions of PCH to the first four excited singlet

states, computed at the CAM-B3LYP/6-311++G(2d,p)/SMD(DCM)//PBE0-GD3BJ/6-31+G(d)/SMD(DCM) level

of theory. The symmetry of PCH was constrained to the D2h point group.

Page 20: Electronic Supporting Information Ultrafast Observation of

S20

Vertical Transition Wavelengths (nm): PCF

Table S1: Vertical transition wavelengths (nm), state-symmetries and oscillator strengths/transition dipole moments

computed for PCF under the C2h point group with TD-DFT (CAM-B3LYP,M06-2X,PBE0-GD3BJ/6-

311++G(2d,p)/SMD(DCM)//PBE0-GD3BJ/6-31+G(d)/SMD(DCM)) and SA-CASSCF/CASPT2 (6-311G(d,p);

CAS(14/14), CAS(8/8)). Oscillator strengths (f) are shown for TD-DFT; transition dipole moments (Debye) are

shown for the SA-CASSCF transitions.

Vertical Transition Wavelengths (nm): PCH

State CAM-B3LYP CAS(14/14)/CASPT2 CAS(8/8)/CASPT2

S1 376 (B1g) 0.00 361 (B1g) 0.00 376 (B1g) 0.00

S2 333 (B2u) 0.206 342(B2u) 4.40 353 (B2u) 4.40

S3 317 (Au) 0.00 310 (Au) 0.00 334 (Au) 0.00

S4 314 (B3g) 0.00 305 (B3g) 0.00 327 (B3g) 0.00

Table S2: Vertical transition wavelengths (nm), state-symmetries and oscillator strengths/transition dipole moments

computed for PCH under the D2h point group with TD-DFT (CAM-B3LYP/6-311++G(2d,p)/SMD(DCM)//PBE0-

GD3BJ/6-31+G(d)/SMD(DCM)) and SA-CASSCF/CASPT2 (6-311G(d,p); CAS(14/14), CAS(8/8)). Oscillator

strengths (f) are shown for TD-DFT; transition dipole moments (Debye) are shown for the SA-CASSCF transitions.

State CAM-B3LYP M06-2X PBE0 CAS(14/14)/

CASPT2

CAS(8/8)/

CASPT2

S1 374 (Ag) 0.00 379 (Ag) 0.00 448 (Au) 0.00 374 (Au) 0.00 380 (Au) 0.00

S2 355 (Au) 0.00 364 (Au) 0.00 443 (Bg) 0.00 366 (Bg) 0.00 376 (Ag) 0.00

S3 352 (Bg) 0.00 360 (Bg) 0.00 418 (Ag) 0.00 362 (Ag) 0.00 374 (Bg) 0.00

S4 337 (Bu) (0.154) 341 (Bu) (0.131) 399 (Bu) (0.042) 353 (Bu) 1.90 371 (Bu) 4.20

Page 21: Electronic Supporting Information Ultrafast Observation of

S21

Scaled Harmonic Frequencies

Transient species Predicted vibrational absorptions / cm-1

1PCF(S0)

1601 (306), 1477 (1494), 1352 (499), 1337 (478)

2PCF

.+(D1) 1603 (160), 1551 (243), 1333 (232)

2MP

.(D1) 1659 (357)

Table S3: Key vibrational absorption frequencies (cm-1

) and corresponding intensities (bracketed) for the ground-

state 1PCF(S0), radical cation

2PCF

.+(D1), and the

2MP

.(D1) radical, computed at the PBE0-GD3BJ/6-

31+G(d)/SMD(DCM) level. The wavenumber values were linearly scaled by a factor of fs = 0.957 ±0.0016 and

used to help assign 2PCF

.+(D1) and

2MP

.(D1) in Figure 1.

Page 22: Electronic Supporting Information Ultrafast Observation of

S22

Reactivity of Debrominated Radical 2MP

.(D1)

B2PLYPD3/6-311++G(2d,p)/SCRF=(SMD,DCM)//PBE0-GD3BJ/6-31+G(d)/ SCRF=(SMD,DCM)

Figure S10: Optimised transition state structures and free energies of activation ∆G‡ for five possible decay

mechanisms of the radical 2MP

.(D1), including: (a) H-atom abstraction from DCM; (b) Cl-atom abstraction from

DCM; (c) H-atom abstraction from MBP; (d) Br-atom abstraction from MBP; and (e) radical addition to the

terminal carbon of methyl methacrylate (MMA). Optimised geometries and thermodynamic corrections were

obtained at the PBE0-GD3BJ/6-31+G(d)/SCRF=(SMD,DCM) level of theory; single-point potential energies

were then computed at the B2PLYPD3/6-311++G(2d,p)/SCRF=(SMD,DCM) level, using the previously

optimised structures. All abstraction reactions have free energies of activation in excess of 80 kJ mol-1

, suggesting

that the radical will be metastable on the picosecond timescale (as observed). Although the reaction of 2MP

.(D1)

with MMA is known to be facile under ambient conditions, no reaction was observed on our timescale; this is

consistent with the computed free energy of activation (52 kJ mol-1

), as for activation-controlled reactions to be

observable on our timescale the barrier should be less than ~ 25 kJ mol-1

.

Page 23: Electronic Supporting Information Ultrafast Observation of

S23

Transient Absorption Spectra

Example Spectral Decompositions

Transient electronic and vibrational spectra were decomposed into constituent parts using the KOALA

program,22

or Origin9 software to identify transient gain features and obtain kinetic traces. Examples of

spectral decompositions are provided in Figure S11 for TEA spectra of PCF in DCM solution (Figure

S1), in Figure S12 for TVA spectra of PCF / MBP in DCM solution (Figure 1(c)), and in Figure S13 for

TEA spectra of PCH or PCF / MBP in DCM solution (Figure 4).

Figure S11: Decompositions of TEA spectra for PCF (2.1 mmol dm-3

) in DCM (Figure S1) at 0.1 ps, 0.2 ps, 0.3

ps, 0.5 ps and 1.0 ps. The fitting residuals at around 430 nm indicate the intermediate electronic states, vibrational

relaxation, or solvation dynamics mentioned in main article. The two basis functions for the decomposition are

shown in the top panel.

0

5

10

15

0

5

10

15

400 450 500 550 600

0

5

10

15

Wavelength / nm

0

5

10

15

Absorbance / m

OD

0

5

10

15

0

5

10

15

TEA spectra

0.1 ps

0.2 ps

0.3 ps

0.5 ps

1.0 ps

5.0 ps0.0 ps

1PCF*(Sn)

TEA spectrum

Total fitting

1PCF*(S1)

10

15

Absorbance / mOD

Page 24: Electronic Supporting Information Ultrafast Observation of

S24

Figure S12: Decompositions of TVA spectra for PCF (2.1 mmol dm-3

) and MBP (0.9 mol dm-3

) in DCM (Figure

1(c) and (d)) at 30 ps, 100 ps, 300 ps and 700 ps. The two basis functions for the decomposition are shown in the

top panel: the TVA spectrum at a time delay of 20 ps was employed as a basis function for the 1PCF*(S1) state,

and that at 1000 ps for the 2PCF・

+ (D1) state. The validity of the latter assignment is confirmed by comparison

with the steady state FT-IR spectrum of 2PCF・

+(D1) obtained by the chemical oxidation method shown in Figure

S5.

0.0

0.2

0.4

Absorbance / mOD

0.0

0.2

0.4

Absorbance / mOD

1530 1540 1550 1560 1570 1580

0.0

0.2

0.4

Wavenumber / cm-1

0.0

0.2

0.4

Absorbance / mOD

0.0

0.2

0.4Absorbance / mOD

TVA spectra

30 ps

100 ps

300 ps

700 ps

1000 ps

20 ps

1PCF*(S1)

2PCF・+(D1)

TVA spectrum

Total fitting

Absorbance / mOD

Page 25: Electronic Supporting Information Ultrafast Observation of

S25

Figure S13: Decompositions of TEA spectra for: (left panel) PCH (2.1 mmol dm-3

) and MBP (1.8 mol dm-3

) in

DCM (Figure 4) at 5 ps, 10 ps, 25 ps and 50 ps; and (right panel) PCF (2.1 mmol dm-3

) and MBP (1.8 mol dm-3

)

in DCM (Figure 4) at 30 ps, 100 ps, 200 ps and 400 ps. The basis functions for decomposition are shown in

Figure S14.

30 ps

100 ps

200 ps

400 ps

5 ps

10 ps

25 ps

50 ps

1PCF*(S1)

2PCF・+(D1)

TEA spectrum

Total fitting

2PCH・+(D1)

1PCH*(S1)

Absorbance / mOD

400 450 500 550 600

0

5

10

15

Wavenumber / cm-1

0

5

10

15

0

5

10

15

0

5

10

15

0

10

20

0

10

20

0

10

20

400 450 500 550 600

0

10

20

Absorbance / mOD

Wavenumber / cm-1

Page 26: Electronic Supporting Information Ultrafast Observation of

S26

TEAS Basis Functions

Figure S14: Individual basis functions used to decompose the TEA spectra in Figure 4. The absorption spectra at

2 ps were taken as the absorption signatures of 1PCH*(S1) (left) or

1PCF*(S1) (right), and the spectra at 1200 ps

as those of the corresponding radical cations.

400 450 500 550 6000

10

20Absorbance / mOD

400 450 500 550 6000

10

20

Absorbance / mOD

400 450 500 550 6000

10

20

Absorbance / mOD

Wavelength / nm

PCH / Toluene

PCH / DCM

PCH / DMF

400 450 500 550 6000

10

400 450 500 550 6000

10

20

400 450 500 550 6000

10

Wavelength / nm

PCF / DMF

PCF / DCM

PCF / Toluene

1PCH*(S1) (2 ps)

2PCH.+(D1) (1200 ps)

1PCF*(S1) (2 ps)

2PCF.+(D1) (1200 ps)

1PCF*(S1) (2 ps)

1PCH*(S1) (2 ps)

1PCF*(S1) (2 ps)1PCH*(S1) (2 ps)

2PCF.+(D1) (1200 ps)

2PCH.+(D1) (1200 ps)

2PCH.+(D1) (1200 ps)

2PCF.+(D1) (1200 ps)

Page 27: Electronic Supporting Information Ultrafast Observation of

S27

Solvent Effects: TEAS Kinetic Traces

Figure S15: Kinetic traces for the TEA spectra in Figure 4 ([PCs]: 2.1 mmol dm-3

, [MBP]: 1.8 mol dm-3

). The

pseudo first-order time constants for PET τ'PET were obtained by fitting to single exponential functions for PCF

and biexponential functions for PCH (see main text). The time constants shown in the figure are the average

values for the decays of 1PC*(S1) state and the rise of

2PC

·+(D1) state. In the case of PCF/MBP in DMF the

observed lifetime was τobs = 153 ps; including the lifetime of 1PCF*(S1) in the absence of MBP, τ1 = 677 ps,

affords a time constant for the PET of τ’PET = 198 ps. In all other cases the S1 lifetime is significantly longer than

τobs, so τobs ~ τ'PET.

PCH / DMF: τʹPET1 = 5.6 ±0.6 ps

PCH / DCM: τʹPET1 = 7.0 ±2.2 ps

PCH / Toluene: τʹPET1 = 3.2 ±0.4 ps

PCF / DMF: τʹPET = 198 ± 7 ps

PCF / DCM: τʹPET = 170 ± 7 ps

PCF / Toluene: τʹPET = 135 ± 7 ps

2PCH・+(D1)

1PCH*(S1)

2PCH・+(D1)

1PCH*(S1)

2PCH・+(D1)

1PCH*(S1)

2PCF・+(D1)

1PCF*(S1)

2PCF・+(D1)

1PCF*(S1)

2PCF・+(D1)

1PCF*(S1)

τʹPET2 = 28 ±3 ps

τʹPET2 = 28 ±4 ps

τʹPET2 = 23 ±3 ps

0.0

0.2

0.4

0.6

0.8

1.0

Integrated intensity

0 50 100 150 200

0.0

0.2

0.4

0.6

0.8

1.0

Integrated intensity

Time / ps

0.0

0.2

0.4

0.6

0.8

1.0

Integrated intensity

0.0

0.2

0.4

0.6

0.8

1.0

Integrated intensity

0 200 400 600 800 1000

0.0

0.2

0.4

0.6

0.8

1.0

Time / ps

0.0

0.2

0.4

0.6

0.8

1.0

Integrated intensity

Page 28: Electronic Supporting Information Ultrafast Observation of

S28

Solvent Effects: Semi-logarithmic Plots of TEAS Kinetic Traces

Figure S16: Semi-logarithmic plots for Figure S15. The plots for PCH show non-single exponential behavior,

whereas those for PCF are linear.

1PCH*(S1)

2PCH・+(D1)2PCF・+(D1)

1PCF*(S1)

0 20 40 60 80 100-4

-3

-2

-1

0

DCM

DMF

toluene

log (intensity)

Time / ps

0 100 200 300 400 500-4

-3

-2

-1

0

DCM

DMF

toluene

log (intensity)

Time / ps

0 20 40 60 80 100-4

-3

-2

-1

0

DCM

DMF

toluene

log (1 - intensity)

Time / ps

0 100 200 300 400 500-4

-3

-2

-1

0

DCM

DMF

toluenelog (1 - intensity)

Time / ps

Page 29: Electronic Supporting Information Ultrafast Observation of

S29

Semi-logarithmic Plots for TVAS Measurements in DCM Solutions

Figure S17: Semi-logarithmic plots for the data presented in Figure 2.

PCH and MBP in DCM: TVA Spectra

Figure S18: TVA spectra for: (a) PCH (2.1 mmol dm-3

) in DCM; and (b) PCH (2.1 mmol dm-3

) and MBP (0.5

mol dm-3

) in DCM. Over 1530 – 1680 cm-1

there is significant spectral congestion, with vibrational absorptions

from at least three normal modes of 1PCH*(S1), two modes for

1PCH(S0) and, in the presence of MBP, from

2PCH

.+(D1) (Figure S6). Although this region was not decomposed successfully, the transient absorption from

2MP

.(D1) at 1660 cm

-1 in the presence of MBP was well-defined and therefore used to measure the rate of the

0 50 100 150 200-3

-2

-1

0

log (1 - intensity)

Time / ps

0 200 400 600 800 1000-3

-2

-1

0

log (1 - intensity)

Time / ps

(a) PCF (b) PCH

(a) PCH

(b) PCH + MBP2MP・(D1)

1550 1600 1650

-0.2

0.0

0.2

Wavenumber / cm-1

-0.2

0.0

0.2

Time / ps

1

10

50

200

500

Absorbance / mOD

Page 30: Electronic Supporting Information Ultrafast Observation of

S30

PET.

Page 31: Electronic Supporting Information Ultrafast Observation of

S31

PCF in DCM: TEAS Concentration Dependence

Figure S19: Kinetic traces for 1PCF*(S1) at various MBP concentrations with TEAS for 2.1 mmol dm

-3 PCF in

DCM solution, with pulsed photoexcitation at 370 nm, and pseudo first-order kinetic plot (inset), affording a

bimolecular rate coefficient of kPET = (4.4 ± 0.3) x 109 dm

3 mol

-1 s

-1. The kinetic traces were obtained using the

same method described in Figure S14.

1.2 1.8 2.4

0.005

0.010

k' PET / s-1

[MBP] / mol dm-3

1PCF*(S1)

0 200 400 600 800 1000

0.0

0.2

0.4

0.6

0.8

1.0

Integrated signal

Time / ps

Page 32: Electronic Supporting Information Ultrafast Observation of

S32

PCH: TEA Spectra

Figure S20: Transient electronic absorption spectra of PCH (2.1 mmol dm-3

) in (a) DMF; and (b) DCM,

following pulsed photoexcitation at 370 nm. In both solvents the decay of the 1PCH*(S1) state is accompanied by

the concomitant formation of a new transient species, assigned to 3PCH*(T1), which grows on too long a

timescale to determine a time constant. The new feature is present at the same position in both solvents and is

inconsistent with the absorption signature of the radical cation 2PCH

.+(D1); taken together, these observations

argue against any suggestion that 1PCH*(S1) decay may be due to electron transfer to the solvent. The lifetime of

the 1PCH*(S1) state is longest in DMF and shortest in DCM, but in both cases the time constant exceeds the

temporal range of our experiment (nanosecond timescale). The formation of a long-lived triplet state 3PCH*(T1) –

which might persist in more dilute solutions of MBP – may be one of the factors that makes PCH a poorer

photocatalyst for controlling polymer molecular weight and polydispersity in O-ATRP. Thermodynamically an

electron transfer from the first excited triplet state is strongly favored, as the 3PCH*(T1) state is not a great deal

lower in energy than the 1PCH*(S1) state and electron transfer from the latter state is seen to be diffusion-

controlled.

(a) DMF (b) DCM1PCH*(S1) 1PCH*(S1)

3PCH*(T1)

3PCH*(T1)

400 450 500 550 6000

10

20

30

Absorbance / m

OD

Wavelength / nm

Time / ps

10

200

300

500

800

1300

400 450 500 550 6000

10

20

30

Wavelength / nm

Page 33: Electronic Supporting Information Ultrafast Observation of

S33

Pure DCM: TEA spectra

Figure S21: Transient electronic absorption spectra of pure DCM following 370-nm excitation. The region at

around 370 nm is masked by the pump light.

Emission Spectra: PCF with/without Triplet Quenchers

Figure S22: Emission spectra for PCF in DCM with and without triplet quencher; pure DCM (∎), styrene (∎),

2,5-dimethylhexa-2,4-diene (DHD,∎), and cyclohexa-1-3-diene (CHD, ∎) following excitation at 370 nm. The

concentration of each triplet quenchers is 1 mol dm-3

. The apparent emission observed at around 450 nm for the

solution containing CHD is most likely due to subtle impurity in CHD.

400 450 500 550 600-5

0

5

10

15

Absorbance / mOD

Wavelength / nm

Time / ps

-0.20

0.00

0.05

0.10

0.50

1.00

2.00

(a)

450 500 550 600 650 7000

50

100

150

200

250

300

Emission intensity / arb. units

Wavelength / nm

without

quencher

styrene

DHD

CHD

(a)

Page 34: Electronic Supporting Information Ultrafast Observation of

S34

Triplet sensitization: PCF in DMF with benzophenone

A triplet sensitizer was added to the PCF solutions in DMF to explore the possible production of the

PCF(T1) triplet excited state using TEAS. The chosen sensitiser was benzophenone (Bzp) because its

triplet state lies sufficiently high in energy to sensitize the excitation of PCF(S0) to PCF(T1) by

intermolecular energy transfer. Figure S23 shows the resulting TEA spectra, and Figure S24

demonstrates that these TEA spectra can be decomposed into features attributed to PCF(S1) and Bzp(T1),

without having to invoke any additional spectral contributions – for example, from PCF(T1).

Figure S23: TEA spectra of a solution of PCF (2.1 mM) and benzophenone

(1.0 M) in DMF. The inset color key shows the time delay at which each

transient spectrum was acquired.

Page 35: Electronic Supporting Information Ultrafast Observation of

S35

Figure S24: Decomposition of selected TEA spectra of a solution of PCF (2.1 mM) and

benzophenone (1.0 M) in DMF into spectral components. The TEA spectra shown were

recorded at time delays of: (a) 30 ps; (b) 50 ps; (c) 100 ps; (d) 200 ps. The spectra used for

the decomposition are shown in panel (e) and were obtained from separate measurements

using PCF/DMF and Bzp/DMF solutions. The accelerated decay of the intensity of the PCF

band in the presence of Bzp is plotted in panel (f).

Steady-state emission measurements of a solution of 0.05 mM PCF and 2.5 mM benzophenone in DMF

showed no significant changes to the intensity of the PCF emission band compared to a 0.05 mM PCF

solution in DMF without added benzophenone. These measurements were made in a 1-cm quartz cuvette,

with an excitation wavelength of 370 nm at which both PCF and benzophenone absorb. This comparison

supports our assignment of the emission to fluorescence from PCF(S1), because phosphorescence from

the PCF(T1) state is expected to be enhanced by benzophenone acting as a triplet sensitizer.

Page 36: Electronic Supporting Information Ultrafast Observation of

S36

References

1. Roberts, G. M.; Marroux, H. J. B.; Grubb, M. P.; Ashfold, M. N. R.; Orr-Ewing, A. J., J. Phys.

Chem. A 2014, 118, 11211-11225.

2. Theriot, J. C.; Lim, C. H.; Yang, H.; Ryan, M. D.; Musgrave, C. B.; Miyake, G. M., Science 2016,

352, 1082-1086.

3. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.;

Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.;

Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.;

Vreven, T.; Montgomery, J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E.;

Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant,

J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J.

B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G.

A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz,

J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Gaussian Inc., Wallingford CT, 2009.

4. Adamo, C.; Barone, V., J. Chem. Phys. 1999, 110, 6158-6170.

5. Marenich, A. V.; Cramer, C. J.; Truhlar, D. G., J Phys Chem B 2009, 113, 6378-6396.

6. Miertus, S.; Scrocco, E.; Tomasi, J., Chem Phys 1981, 55, 117-129.

7. Tomasi, J.; Mennucci, B.; Cammi, R., Chem. Rev. 2005, 105, 2999-3093.

8. Grimme, S.; Ehrlich, S.; Goerigk, L., J. Comput. Chem. 2011, 32, 1456-1465.

9. Goerigk, L.; Grimme, S., J. Chem. Theory Comput. 2011, 7, 291-309.

10. Essafi, S.; Tomasi, S.; Aggarwal, V. K.; Harvey, J. N., J. Org. Chem. 2014, 79, 12148-12158.

11. Adamo, C.; Jacquemin, D., Chem Soc Rev 2013, 42, 845-856.

12. Head-Gordon, M.; Rico, R. J.; Oumi, M.; Lee, T. J., Chem. Phys. Lett. 1994, 219, 21-29.

13. Laurent, A. D.; Jacquemin, D., Int. J. Quantum Chem. 2013, 113, 2019-2039.

14. Jacquemin, D.; Perpete, E. A.; Scuseria, G. E.; Ciofini, I.; Adamo, C., J. Chem. Theory Comput.

2008, 4, 123-135.

15. Peach, M. J. G.; Benfield, P.; Helgaker, T.; Tozer, D. J., J. Chem. Phys. 2008, 128, 044118

16. Yanai, T.; Tew, D. P.; Handy, N. C., Chem. Phys. Lett. 2004, 393, 51-57.

17. Werner, H.-J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz, M.; Celani, P.; Korona, T.; Lindh,

R.; Mitrushenkov, A.; Rauhut, G.; Shamasundar, K. R.; Adler, T. B.; Amos, R. D.; Bernhardsson,

A.; Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Goll, E.; Hampel, C.;

Hesselmann, A.; Hetzer, G.; Hrenar, T.; Jansen, G.; Köppl, C.; Liu, Y.; Lloyd, A. W.; Mata, R. A.;

May, A. J.; McNicholas, S. J.; Meyer, W.; Mura, M. E.; Nicklass, A.; O'Neill, D. P.; Palmieri, P.;

Pflüger, K.; Pitzer, R.; Reiher, M.; Shiozaki, T.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson,

T.; Wang, M.; Wolf, A., MOLPRO, version 2010.1, a package of ab initio programs. see

http://www.molpro.net.

18. Werner, H. J.; Knowles, P. J., J. Chem. Phys. 1985, 82, 5053-5063.

19. Knowles, P. J.; Werner, H. J., Chem. Phys. Lett. 1985, 115, 259-267.

20. Silva, M. R.; Schreiber, M.; Sauer, S. P. A.; Thiel, W., J. Chem. Phys. 2010, 133, 174318.

21. Guido, C. A.; Knecht, S.; Kongsted, J.; Mennucci, B., J. Chem. Theory Comput. 2013, 9, 2209-2220.

22. Grubb, M. P.; Orr-Ewing, A. J.; Ashfold, M. N. R., Rev. Sci. Instrumen. 2014, 85, 064104.