Fact-checked by Grok 2 weeks ago

Time-resolved spectroscopy

Time-resolved spectroscopy encompasses a suite of analytical techniques that measure the temporal evolution of a material's optical, , or structural properties following an event, enabling the observation of ultrafast dynamic processes on timescales ranging from femtoseconds to microseconds. These methods typically rely on sources to initiate and probe transient states, providing insights into molecular motions, energy transfer, and reaction pathways that are inaccessible through steady-state . The foundational principle of time-resolved spectroscopy is the pump-probe methodology, where a "pump" excites the sample to a non-equilibrium , and a subsequent "probe" interrogates the system's response after a controlled delay, often as short as 10 femtoseconds (10^{-14} s). This approach, rooted in and the time-dependent , allows for the resolution of atomic-scale dynamics, such as vibrational relaxation, , and bond breaking/forming, by tracking changes in , , , or spectra. Key variants include time-resolved photoelectron spectroscopy (TRPES) for electronic mapping, resonance Raman for vibrational signatures, and for structural evolution, with determined by duration and synchronization precision. Historical milestones trace back to the with by Norrish and Porter, but the field's transformative leap occurred in the 1980s through innovations, earning Ahmed the 1999 for establishing femtochemistry—a subfield focused on real-time dynamics. Applications of time-resolved spectroscopy span , physics, , and , revealing mechanisms in photochemical reactions, photosynthetic energy transfer, and charge dynamics. For instance, it has elucidated isomerization in (∼200 fs) and charge separation in solar cells, informing the design of efficient photovoltaic devices. In , it probes and , while in condensed matter, it studies phase transitions and spin dynamics in perovskites. Ongoing advancements, such as extensions and integration, continue to push resolution limits, bridging atomic motions with macroscopic phenomena.

Fundamentals

Definition and Principles

Time-resolved spectroscopy is the study of transient changes in the , , or spectra of following an event, enabling the of dynamic processes on ultrafast timescales. This approach typically employs pulsed light sources to achieve high , allowing researchers to capture the of molecular or states after . Unlike steady-state , which measures time-averaged properties under continuous illumination and reflects conditions, time-resolved methods resolve non-equilibrium by correlating spectral changes with precise time delays. The core principle involves time correlation between an (often called the ) and a subsequent probe , where the delay between them, denoted as τ, is systematically varied to map out temporal evolution. In pump-probe setups, the induces a in the sample, and the probe interrogates its spectral response, yielding a signal that depends on τ. The measured signal S(τ) is fundamentally a of the instrument response function—limited primarily by the duration—with the sample's dynamic response, accounting for finite in the . A basic expression for the pump-probe signal in is given by S(\tau) = \int \epsilon(\lambda, t) \cdot I_{\text{probe}}(\lambda, t - \tau) \, dt, where \epsilon(\lambda, t) is the time-dependent extinction coefficient at wavelength \lambda, and I_{\text{probe}}(\lambda, t - \tau) is the probe intensity delayed by τ; this integral highlights how the signal integrates the evolving optical properties over the probe's temporal profile. These techniques span a wide range of timescales, from femtoseconds (10^{-15} s) for molecular vibrations and electronic transitions to picoseconds (10^{-12} s) and milliseconds (10^{-3} s) for relaxation processes such as vibrational cooling and non-radiative decay. For instance, femtosecond resolutions probe ultrafast electronic excitations and initial vibrational coherences, while longer scales capture thermalization and population transfers in condensed phases. This temporal coverage provides insight into fundamental mechanisms like energy transfer and reaction pathways that are averaged out in steady-state measurements.

Historical Development

The origins of time-resolved spectroscopy trace back to the early , when researchers sought to capture the temporal evolution of light from transient sources. In 1900, Arthur Schuster and G. Hemsalech conducted the first reported experiments on time-resolved optical spectroscopy, employing a technique that involved moving in the focal plane of a spectrograph to record the duration and sequence of lines in electric . This approach allowed rudimentary of emission lifetimes on the order of microseconds, laying foundational groundwork for studying dynamic processes in luminous discharges. Significant advancements occurred in the mid-20th century with the development of , a method pioneered by Ronald G. W. Norrish and in 1949. This technique utilized high-intensity flash lamps to initiate rapid photochemical reactions, followed by spectroscopic detection of short-lived intermediates, achieving time resolutions down to microseconds. Their work enabled the direct observation of free radicals and transient species in gas-phase reactions, revolutionizing the study of fast and earning them, along with , the 1967 for studies of extremely fast chemical reactions. The 1960 invention of the by further transformed the field, introducing coherent, pulsed light sources that facilitated precise excitation and probing in time-resolved experiments during the 1960s, marking the shift toward laser-based transient spectroscopy. The 1970s brought picosecond time resolution through innovations in mode-locked dye lasers and detection methods. Erich P. Ippen and Charles V. Shank demonstrated subpicosecond spectroscopy using colliding-pulse mode-locking in dye lasers, enabling the study of ultrafast relaxation processes in condensed phases with pulses as short as 100 femtoseconds by the decade's end. Concurrently, the development of streak cameras in the early 1970s, such as those by David J. Bradley and Wilson Sibbett, provided electronic time-resolved imaging with resolutions below 10 picoseconds, essential for capturing transient spectra in real time. The ultrafast era accelerated in the 1980s with the advent of lasers, exemplified by R. L. Fork and colleagues' 1982 demonstration of to 30 using and grating-pair dispersion compensation in dye systems. This enabled Ahmed H. Zewail's pioneering experiments starting in 1987, where pump-probe techniques resolved atomic motions during bond breaking and formation on timescales, culminating in his 1999 . The 1990s saw the commercialization of titanium-sapphire (Ti:sapphire) lasers, first demonstrated by Peter F. Moulton in 1982 and widely available by 1990 through companies like Spectra-Physics, providing tunable, high-repetition-rate pulses that became standard for ultrafast . Progression to scales occurred in the , with the first isolated pulses generated in 2001 via high-harmonic generation (HHG) by Michael Hentschel and colleagues, achieving a pulse duration of 650 for dynamics. Post-2010 advances in HHG, including improved phase-matching and ionization control in gas jets, have enhanced pulse isolation and flux, enabling applications in core-level and tracking. In 2023, , , and were awarded the for their pioneering experimental methods that generate pulses of light for the study of dynamics in matter.

Instrumentation

Ultrafast Light Sources

Ultrafast sources are critical for time-resolved spectroscopy, providing short-duration pulses that enable the of molecular and on femtosecond to picosecond timescales. These sources typically generate coherent pulses with durations ranging from () to (), allowing comparable to the processes under study. Mode-locked lasers form the backbone of such systems, where the locking of longitudinal cavity modes produces a train of ultrashort pulses at repetition rates from kilohertz (kHz) to megahertz (MHz). A prominent example is the titanium-doped sapphire (Ti:sapphire) oscillator, which operates around 800 nm and routinely produces pulses of approximately 100 fs duration with average powers exceeding 1 W at 80 MHz repetition rates. These oscillators achieve such short pulses through Kerr-lens mode-locking, a self-focusing mechanism that favors the lowest-intensity parts of the pulse to build up ultrashort durations without external modulators. For applications requiring higher pulse energies, amplified systems employ , a technique that stretches the pulse temporally before to avoid optical damage, then compresses it back to near its original duration, yielding millijoule-level energies while maintaining widths. Key performance parameters include pulse duration, repetition rate, and wavelength tunability, which dictate the source's suitability for specific spectroscopic probes. The minimum achievable pulse duration is governed by the time-bandwidth product, a fundamental limit from relations; for Gaussian pulses, it is expressed as \Delta t \approx \frac{0.44}{\Delta \nu}, where \Delta t is the pulse duration (full width at half maximum, FWHM) and \Delta \nu is the spectral bandwidth in frequency units, highlighting the trade-off between temporal shortness and spectral breadth. Wavelength tunability is often provided by optical parametric amplifiers (OPAs), which convert the output of a into broadly tunable pulses from the visible to mid-infrared (e.g., 400–2500 nm) via nonlinear parametric down-conversion in crystals like beta-barium borate, enabling excitation or probing across diverse electronic and vibrational transitions. Beyond Ti:sapphire-based systems, other generation methods expand the spectral and temporal reach. fiber lasers, often ytterbium- or erbium-doped, offer compact alternatives with pulse durations of 100–500 fs and repetition rates up to 100 MHz, suitable for integration into portable setups. Supercontinuum sources, generated by launching pulses into fibers or bulk materials, produce broadband white-light continua spanning hundreds of nanometers (e.g., 450–2000 nm) through and effects, providing versatile probe beams for spectroscopy. For (XUV) and regimes, high-harmonic generation (HHG) drives intense pulses into a gaseous medium to produce coherent harmonics up to keV energies, yielding isolated pulses (e.g., 25 as duration) for probing sub-femtosecond dynamics. In pump-probe configurations, precise synchronization between (pump) and (probe) pulses is essential for accurate time-resolved measurements, typically achieved with optical delay lines that adjust path lengths on motorized stages for sub-femtosecond control, or electronic triggers for high-repetition-rate locking. Stable timing, often below 100 RMS, ensures reliable mapping of transient signals without artifacts from pulse-to-pulse variations.

Detection and Time-Resolution Methods

In time-resolved spectroscopy, detection methods capture the temporal evolution of spectral signals following excitation by pulsed light sources. Common detectors include (CCD) cameras, which provide high for spectrally dispersed signals across multiple wavelengths simultaneously, enabling acquisition in setups like transient absorption experiments. Photomultiplier tubes (PMTs) offer single-photon sensitivity and fast response times, often used in time-correlated single-photon counting (TCSPC) configurations to measure decays with high temporal precision. Streak cameras achieve direct time resolution by converting temporal signal variations into spatial distributions via electron deflection, allowing to imaging of ultrafast events without mechanical scanning. For in continuous or chopped signal schemes, lock-in amplifiers demodulate the signal at a known reference , suppressing noise and improving detection limits by orders of magnitude in low-light conditions. Time-resolution techniques synchronize or encode the arrival times of excitation and probe signals to map dynamics. Optical delay lines, consisting of motorized translation stages, scan the probe pulse relative to the excitation to sample kinetics point-by-point, achieving resolutions limited by stage precision and pulse durations. Asynchronous methods, such as optical Kerr gating, employ a nonlinear medium where a gate pulse induces transient birefringence to "slice" the signal with femtosecond precision, bypassing mechanical delays for broadband detection. Upconversion techniques mix the time-varying signal with a delayed gate pulse in a second-order nonlinear crystal via sum-frequency generation, converting the signal to a higher-frequency detectable range while providing sub-picosecond gating. Spatial encoding via asynchronous optical sampling (ASOPS) uses two mode-locked lasers with slightly offset repetition rates to rapidly scan time delays electronically, enabling high-repetition-rate measurements over nanosecond windows without moving parts. The ultimate is constrained by the instrument response function (IRF), which the intrinsic sample dynamics with the system's response, broadening observed features. For ultrafast setups, the IRF typically ranges from 50 to 100 fs, determined by durations and detector . The effective follows Gaussian convolution, where the total standard deviation is given by \sigma_{\text{total}} = \sqrt{\sigma_{\text{sample}}^2 + \sigma_{\text{IRF}}^2}, quantifying how system limitations degrade the sample's native timescale. Data handling in time-resolved spectroscopy emphasizes optimizing signal quality and extracting kinetics. (SNR) is enhanced by accumulating multiple scans and averaging, with considerations for dominance at low intensities and 1/f noise at longer delays. Background subtraction removes or contributions, often via pump-off spectra to isolate signals. Global fitting analyzes multi-wavelength datasets simultaneously using shared kinetic models, improving parameter accuracy for species-associated spectra and rate constants compared to wavelength-independent fits.

Absorption and Emission Techniques

Transient Absorption Spectroscopy

Transient absorption spectroscopy is a pump-probe technique that monitors the transient changes in the absorption spectrum of a sample following photoexcitation, providing insights into the dynamics of excited states on to timescales. In this method, a excites a fraction of the molecules in the sample to an excited state, while a time-delayed probe pulse interrogates the resulting changes in optical density across a broad spectral range. The differential absorbance, denoted as \Delta A(\lambda, \tau), is calculated as \Delta A(\lambda, \tau) = -\log_{10} \left( \frac{I_{\text{probe}}(\lambda, \tau)}{I_{\text{probe0}}(\lambda)} \right), where I_{\text{probe}}(\lambda, \tau) is the probe intensity after excitation at delay time \tau, and I_{\text{probe0}}(\lambda) is the unexcited reference intensity. This yields a two-dimensional map of \Delta A as a function of wavelength \lambda and time delay \tau, revealing key photophysical processes. The experimental setup typically involves a femtosecond system, such as a Ti: oscillator and , generating pulses at wavelengths tuned to the sample's (often 400-800 nm) with energies of 5-100 nJ and repetition rates of 1-250 kHz to minimize heating effects. The probe is a broadband white-light continuum, generated by focusing a portion of the laser output into a or _2 crystal, spanning the UV-Vis to near-infrared () range (typically 350-750 nm, extendable to 1600 nm). The and probe beams are spatially overlapped in the sample, with the probe delayed relative to the using an optical delay line, and the transmitted probe is dispersed via a spectrometer onto a array or detector for multichannel detection. Spectral features in the \Delta A map include negative signals from ground-state bleach (depletion of the ground-state population) and (from the to the ground state), alongside positive signals from (to higher-lying states). These signatures enable mapping of electronic transitions and relaxation pathways, commonly applied to study charge transfer and in molecular systems. Data analysis begins with preprocessing to correct for artifacts like in the probe and scattered pump light, followed by extraction of kinetic traces at selected wavelengths. Individual traces are often fitted to multi-exponential decay models to quantify time constants associated with processes like vibrational relaxation or , though global analysis across the full dataset is preferred for correlated dynamics. (SVD) serves as a dimensionality reduction step, identifying the number of significant spectral components, which informs subsequent fitting to models such as evolution-associated difference spectra (EADS) or species-associated difference spectra (SADS). Validity requires low excitation densities (e.g., <1% of molecules excited per pulse) to ensure a linear response and avoid nonlinear effects like exciton annihilation or multiphoton processes. Spectral resolution is achieved through grating spectrometers, typically providing 0.1-1 resolution depending on the grating and slit width, sufficient to resolve molecular electronic bands in the UV-Vis-NIR range. is fundamentally limited by the pump-probe temporal overlap and pulse durations, often reaching sub-50 with non-collinear optical parametric amplifiers for broadband pulses, though and can broaden the instrument response function to 100-200 . These resolutions allow precise tracking of ultrafast transients, such as sub-picosecond charge separation in donor-acceptor systems.

Time-Resolved Fluorescence Spectroscopy

Time-resolved fluorescence spectroscopy measures the temporal evolution of emission following , providing insights into excited-state lifetimes, processes, and molecular reorientation on to timescales. This technique relies on the of photons from electronically excited molecules, allowing the probing of dynamic processes that are inaccessible to steady-state measurements. By resolving the and of the emitted , it reveals information about the local microenvironment, such as solvent interactions and biomolecular conformations. The primary methods for time-resolved fluorescence spectroscopy include time-correlated single-photon counting (TCSPC), streak cameras, and fluorescence upconversion. TCSPC is widely used for to decay measurements, where a pulsed source is employed, and the arrival time of individual emission s is recorded relative to the using a time-to-amplitude converter and ; this statistical approach builds a of arrival times to reconstruct the decay profile with high sensitivity and dynamic range up to 10^6. Streak cameras, suitable for to resolutions, convert s to electrons via a photocathode, accelerate them electrostatically, and sweep them across a screen to capture time- and -resolved images of the emission. Fluorescence upconversion enables ultrafast (sub-) measurements by sum-frequency mixing the fluorescence signal with a gate in a nonlinear , shifting it to a detectable for . These methods complement each other, with TCSPC offering the highest signal-to-noise for longer timescales and upconversion providing the finest limited primarily by the duration. Key observables in time-resolved fluorescence spectroscopy include the fluorescence lifetime \tau, derived from the intensity decay modeled as I(t) = I_0 \exp(-t/\tau), which quantifies the average time a molecule spends in the excited state before emission and is sensitive to quenching or energy transfer mechanisms. For rotational dynamics, time-resolved anisotropy r(t) is calculated as r(t) = \frac{I_\parallel(t) - I_\perp(t)}{I_\parallel(t) + 2I_\perp(t)}, where I_\parallel and I_\perp are the parallel and perpendicular polarized intensities, respectively; this decays due to rotational diffusion, with the correlation time \phi reflecting molecular size and viscosity via the Perrin equation \frac{1}{r(t)} = \frac{1}{r_0} \left(1 + \frac{t}{\phi}\right). These parameters allow deconvolution of multi-exponential decays in heterogeneous systems, such as multi-fluorophore environments. Variants of the technique extend its utility to specific dynamics. Time-resolved measurements track orientational relaxation and homo-FRET in oligomeric proteins, where arises from energy transfer between identical fluorophores. Fluorescence resonance energy transfer () dynamics are probed by monitoring donor lifetime shortening in the presence of an acceptor, with transfer efficiency E = 1 - \tau_{DA}/\tau_D, revealing distance changes on the 1-10 scale during biomolecular interactions. shifts due to solvatochromism are captured by time-resolved spectra, where excited-state moments cause red- or blue-shifts in polar solvents, indicating solvation dynamics on to timescales. These approaches are particularly sensitive to the local environment, enabling studies of and conformational changes, though they are inherently limited to intrinsically or extrinsically fluorescent samples.33209-4)

Vibrational Techniques

Time-Gated Raman Spectroscopy

Time-gated Raman spectroscopy employs a pump-probe scheme to investigate vibrational dynamics in photoexcited species, where an ultrafast pump pulse promotes the sample to an excited electronic state, and a subsequent probe pulse generates Raman scattering that is temporally selected using a fast gate to capture signals from transient intermediates. This approach isolates the instantaneous Raman emission, which occurs on femtosecond to picosecond timescales, from slower fluorescence or ground-state contributions, enabling resolution of vibrational modes unique to the excited state, such as altered bond strengths reflected in Stokes (lower energy) and anti-Stokes (higher energy) shifts. In typical setups, the is a to laser pulse, while the is a pulse synchronized with the , achieving time resolutions from picoseconds (using optical Kerr gates) to nanoseconds (with intensified detectors). enhancement of the Raman signal, by tuning the wavelength to match absorptions of the transient , can increase sensitivity by orders of magnitude, facilitating detection of weak excited-state spectra in complex systems like biomolecules or materials. The Raman shift, a measure of the vibrational , is quantified as \Delta \nu = \frac{1}{\lambda_\text{laser}} - \frac{1}{\lambda_\text{scattered}}, where \lambda_\text{laser} is the probe and \lambda_\text{scattered} is the of the scattered , expressed in wavenumbers (cm^{-1}). Analysis involves tracking time-dependent evolutions in peak positions and intensities across delay scans; for instance, coherent vibrational wavepacket motions manifest as oscillatory modulations, while decay reveal intramolecular processes. This technique uniquely differentiates ultrafast intramolecular vibrational redistribution (IVR) on timescales from picosecond vibrational cooling toward equilibrium in the , providing insights into energy flow not accessible via steady-state methods. Its adoption remains limited compared to continuous-wave , primarily due to the challenges in amplifying inherently faint transient signals amid noise.

Time-Resolved Infrared Spectroscopy

Time-resolved (IR) employs ultrafast mid-IR pulses in the 2-20 μm wavelength range to excite and vibrational modes, enabling the observation of transient structural and dynamic changes on to timescales. In a typical - setup, a mid-IR selectively excites a vibrational transition, while a delayed measures the resulting changes in across a broad spectral range. This allows for the direct monitoring of and relaxation processes in molecular systems. For broadband detection, step-scan Fourier-transform () is often utilized, where the interferometer mirror is stepped to specific positions, and time-resolved interferograms are recorded at each step to reconstruct full spectra with resolution or better. The core of time-resolved IR spectroscopy lies in tracking vibrational lifetimes and energy flow through measurements of the transient absorbance change, denoted as ΔA(ν, τ), where ν is the probe frequency and τ is the pump-probe delay. This signal arises from ground-state bleaching, , and excited-state absorption, revealing how vibrational energy redistributes within molecules and their surroundings. In hydrogen-bonded networks, such as those in or biomolecules, energy redistribution occurs rapidly, often on scales, as excess vibrational energy cascades through low-frequency modes, leading to dynamics and structural rearrangements. For instance, excitation of the O-H stretch in results in ultrafast energy transfer to the H-bond network, observable as shifts and broadening in the IR spectrum. Advanced analysis in time-resolved IR spectroscopy frequently involves two-dimensional IR (2D-IR) techniques, which map correlations between excitation and detection frequencies to disentangle homogeneous and inhomogeneous broadening. These frequency-frequency correlations provide insights into coupling between vibrational modes and the timescale of structural fluctuations. Anharmonicity plays a key role in these spectra, causing frequency shifts upon vibrational excitation. A particular strength of time-resolved IR spectroscopy is its sensitivity to carbonyl stretches in proteins, which appear around 1600-1700 cm⁻¹ and serve as site-specific probes for secondary structure and hydrogen bonding. These modes report on conformational changes during folding or enzymatic reactions, with transient signals revealing hydrogen-bond alterations. With resolution, the technique captures intramolecular vibrational redistribution (IVR), where energy from a localized spreads across the molecule's vibrational manifold in tens of femtoseconds, as demonstrated in studies of polyatomic molecules. This enables real-time visualization of how vibrational evolves before dissipation into the solvent.

Photoelectron and Advanced Techniques

Time-Resolved Photoemission Spectroscopy

Time-resolved photoemission spectroscopy (TRPES) employs a - approach to ultrafast electronic processes by exciting a sample with an ultrashort , which populates transient electronic states, followed by a delayed in the (UV) or (XUV) range that ejects photoelectrons from occupied orbitals. These photoelectrons carry information about the energy, momentum, and temporal evolution of the electronic structure, allowing direct observation of charge dynamics and relaxation mechanisms. The of the emitted photoelectrons, E_{\kin}, is governed by E_{\kin} = h\nu - \IP - \phi, where h\nu is the probe , \IP is the ionization potential () of the initial state, and \phi is the sample ; this relation enables precise determination of binding energies at various time delays. Detection typically relies on time-of-flight (TOF) spectrometers, which measure flight times to infer energies, or hemispherical analyzers for combined energy and angular resolution, achieving femtosecond temporal precision. Key variants include angle-resolved TRPES (TR-ARPES), which captures the angular distribution of photoelectrons to resolve momentum-space information and map electronic dispersions, and core-level TRPES, which uses energetic XUV probes to access localized core orbitals for site-specific electronic probing. These extensions enhance the technique's ability to disentangle orbital symmetries and spatial charge distributions. Analysis of TRPES data focuses on time-dependent shifts in binding energies, which signal charge transfer and screening effects, as well as the evolution of photoelectron yields to track femtosecond-scale hot electron dynamics, including thermalization via electron-electron scattering. For instance, binding energy shifts on the order of 0.1–1 eV can indicate interfacial charge redistribution in heterostructures. As a vacuum-based method, TRPES operates under (<10^{-10} mbar) to prevent and maintain surface cleanliness, particularly essential for solid samples where it reveals momentum-resolved band structures and non-equilibrium . With attosecond-duration probes, it achieves sub-femtosecond resolution to observe intrinsic motion, such as field-driven tunneling and recollision processes. These XUV probes are generated via high-harmonic generation from ultrafast lasers.

Two-Photon Photoemission (2PPE)

Two-photon photoemission (2PPE) is a nonlinear optical that probes unoccupied electronic states above the by sequentially absorbing two photons from ultrashort pulses. In the , the first photon excites an from an occupied state, typically below the , to an unoccupied below the level; the second photon then ionizes the electron from this intermediate state into the vacuum, producing a detectable photoelectron. The photoemission yield scales quadratically with the intensity (∝ I²), distinguishing it from single-photon processes, and the intermediate state's lifetime (τ_int) governs the efficiency and of the measurement. The efficiency of the two-photon is proportional to the intermediate state's lifetime τ_int, as the transient population in the intermediate state determines the probability of subsequent ; this reflects the intermediate state's role in bridging the two steps, stemming from perturbative treatments balancing excitation rates and decay. Experimental setups for 2PPE employ pump-probe schemes, where a pump pulse (often in the visible or near-infrared range, e.g., 400–800 nm) populates the , and a delayed probe pulse ionizes it, enabling time-resolved studies on to timescales. Photoelectrons are collected and analyzed using hemispherical electron energy analyzers, which provide both resolution (typically ~10–50 meV) and, in angle-resolved configurations, information parallel to the surface. This setup, as detailed in early implementations, allows mapping of transient electronic structures at surfaces with high temporal precision. Analysis of 2PPE data involves fitting time-delay scans to extract intermediate state lifetimes and dynamics, often revealing quantum beats—coherent oscillations in the signal arising from superposition of multiple s, such as image-potential states on metal surfaces. These beats provide insights into and between states, with periods determined by energy splittings (e.g., ~100–500 for Rydberg-like image states). 2PPE is particularly sensitive to surface and phenomena due to its reliance on low-energy intermediate states localized near the surface, making it ideal for studying image-potential states, adsorbate-induced modifications, and at clean metal surfaces or thin films. For instance, adsorbates like alkali metals on Cu(111) alter state lifetimes via hybridization, detectable as shifts in peak energies or decay rates. In modern applications, 2PPE has been extended to two-dimensional materials, where it probes excitonic states; in black phosphorus, interferometric 2PPE reveals ultrafast of parity-forbidden excitons on ~100 fs scales, highlighting its utility for low-dimensional systems with reduced screening.

Other Multiple-Pulse Techniques

Other multiple-pulse techniques in time-resolved spectroscopy extend beyond simple pump-probe schemes by employing coherent nonlinear interactions among three or more pulses to probe ensemble dynamics with enhanced selectivity. These methods, rooted in (FWM) processes, enable the isolation of specific quantum pathways and the measurement of , , and couplings in complex systems. Transient grating spectroscopy creates a spatial of or via of two excitation pulses, forming a transient diffraction grating that diffracts a third probe pulse to reveal processes. This technique measures translational diffusion coefficients by tracking the grating decay time, which follows an form τ = Λ² / (4D), where Λ is the grating fringe spacing and D is the diffusion constant. In semiconductors, it has quantified diffusion on timescales, providing insights into transport without invasive probes. Photon echo spectroscopy uses a three-pulse to rephase inhomogeneous in ensembles, generating an signal that refocuses the coherent response after a delay. The intensity decays with the pure time T₂*, isolating homogeneous linewidths from static . This method has been applied to molecular crystals, revealing rates on the order of 1-10 in photosynthetic complexes. Two-dimensional (2D) spectroscopy employs three pulses in a non-collinear geometry to map couplings and coherences, resolving excitation and emission frequencies in a spectrum. Off-diagonal peaks indicate excitonic interactions, while diagonal peaks reflect , with resolution capturing quantum beats. In the Fenna-Matthews-Olson (FMO) complex, it has observed long-lived coherences persisting for 660 fs at 77 K, suggesting wavelike energy transfer in . These techniques rely on phase-matching in FWM, where the wavevector of the output signal k_s = -k_1 + k_2 + k_3 ensures efficient nonlinear generation and directional emission. Rephasing pulse sequences, such as those in photon echoes or 2D spectroscopy, conjugate the phase of the first interaction to reverse pathways, yielding absorptive line shapes that enhance signal clarity. The underlying nonlinear response is described by the third-order susceptibility χ^(3), which governs the : \mathbf{P}^{(3)}(t) = \epsilon_0 \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \chi^{(3)}(t - t_3, t_3 - t_2, t_2 - t_1) : \mathbf{E}(t_3) \mathbf{E}(t_2) \mathbf{E}(t_1) \, dt_1 \, dt_2 \, dt_3 This captures time-ordered interactions, with the Liouville pathway isolating rephasing versus non-rephasing contributions. Analysis of these signals produces 2D maps plotting frequency versus coherence time or frequency, revealing energy transfer pathways through cross-peak evolution. In coherent control applications, shaped multiple pulses selectively enhance or suppress reaction branches, such as bond-selective in polyatomic molecules, by interfering pathways with differences.

Applications

In Chemical Dynamics

Time-resolved spectroscopy has revolutionized the study of chemical dynamics by enabling direct observation of transient species and reaction pathways on ultrafast timescales, providing insights into the mechanisms of bond breaking, formation, and energy redistribution in molecular systems. Pioneering experiments, such as those conducted by Ahmed Zewail's group, demonstrated the real-time of the iodine-carbon bond in ICN following photoexcitation, revealing a wavepacket evolution along the with a dissociation time of approximately 200 fs in the gas phase. This work established time-resolved techniques as essential tools for capturing the atomic-scale motions governing unimolecular reactions, highlighting the role of surfaces in directing product formation. In the , advancements extended these observations to bond-forming processes, illustrating concerted mechanisms in polyatomic systems. For instance, femtosecond studies of photoexcited methylene iodide (CH₂I₂) captured the synchronous breaking of two C-I bonds and formation of an I-I bond, occurring on a timescale of less than 100 fs without intermediate species, thus confirming the concerted nature of the reaction in the gas phase. Such experiments underscored the capability of time-resolved methods to resolve multidimensional landscapes, where vibrational drives efficient during recombination. Key processes elucidated include ultrafast and proton transfer, which are central to understanding evolution. In cis-stilbene, time-resolved and tracked the torsional motion leading to trans-stilbene formation via a , with the excited-state lifetime measured at around 70 fs in the gas phase, emphasizing the role of nonadiabatic couplings in facilitating rapid ground-state recovery. Similarly, in the 7-azaindole dimer—a model for base-pair tautomerization— pump-probe observed double proton transfer dynamics, completing in less than 300 fs and revealing barrierless pathways influenced by hydrogen bonding. These studies highlight how time-resolved probes the fleeting s, linking structural changes to reactivity. Transient absorption spectroscopy has been instrumental in mapping conical intersections, where electronic states degeneracy enables ultrafast nonradiative decay, typically on timescales for curve crossings (e.g., 10–100 fs in polyenes). In solution, solvation dynamics introduce picosecond components (1–10 ps) due to diffusive reorganization around the solute, as observed in probe molecules like Coumarin 153, where inertial solvent responses occur in ~50 fs followed by slower relaxation. Collectively, these measurements yield direct quantification of reaction rates and quantum yields—for example, determining the branching ratios in ICN to correlate wavepacket with product distributions—thus providing foundational benchmarks for theoretical models of chemical reactivity.

In Biological Systems

Time-resolved spectroscopy has proven invaluable for elucidating biomolecular dynamics in biological systems, capturing ultrafast processes such as , , and energy transfer within photosynthetic complexes on timescales ranging from femtoseconds to nanoseconds. These techniques reveal transient structural changes and interactions that underpin cellular functions, providing insights into how proteins achieve functional conformations or facilitate chemical reactions . For instance, in , time-resolved X-ray scattering has tracked the redox-coupled folding of , showing heterogeneous unfolded states that evolve into compact structures within microseconds following electron injection. Similarly, in , ultrafast time-resolved spectroscopy has illuminated light-driven mechanisms in enzymes like [FeFe]-hydrogenases, where picosecond dynamics govern steps essential for . A prominent example is the femtosecond study of retinal isomerization in rhodopsin, the light-sensitive protein in vertebrate vision. Upon photon absorption, the 11-cis retinal chromophore undergoes torsional isomerization to the all-trans form in approximately 200 femtoseconds, initiating the visual signal transduction cascade. This ultrafast reaction, captured via femtosecond transient absorption spectroscopy, highlights the protein's role in constraining the chromophore to achieve near-unity quantum efficiency. In photosynthetic complexes, time-resolved fluorescence spectroscopy has mapped excitation energy transfer dynamics, such as in phycobilisomes of cyanobacteria, where energy migrates from phycobiliproteins to photosystem II core antennas on picosecond timescales, optimizing light harvesting under varying conditions. Two-dimensional infrared (2D-IR) spectroscopy has been instrumental in monitoring amyloid fibril formation, a process implicated in diseases like type 2 diabetes. Isotope labeling combined with 2D-IR reveals a multistep pathway for human islet amyloid polypeptide (hIAPP), involving transient β-sheet-rich intermediates during the lag phase before fibril elongation, with spectral signatures evolving over minutes to hours. Electron and proton transfer processes in DNA, critical for replication and repair, occur on femtosecond to picosecond scales; for example, UV-induced interstrand proton transfer, triggered by intrastrand electron transfer, has been observed using femtosecond time-resolved infrared spectroscopy in model duplexes, demonstrating long-lived excited states that facilitate charge migration without permanent damage. Vibrational relaxation in peptides, which dissipates excess energy post-excitation, proceeds rapidly at biological interfaces. Time-resolved studies of I vibrations in membrane-bound α-helical peptides show relaxation times of about 1.7 picoseconds, influenced by the hydrophobic core of lipid bilayers and surrounding . On longer picosecond-to-nanosecond scales, conformational changes in proteins, such as domain rearrangements, are probed by time-resolved of the fluorescence , revealing solvent relaxation and side-chain motions that stabilize folded states. Time-resolved fluorescence also enables (FRET) measurements in protein domains, quantifying intramolecular distances and interactions in live cells with sub-nanosecond resolution, as demonstrated in imaging protein-protein associations. Despite these advances, applying time-resolved spectroscopy to biological systems faces challenges, including sample heterogeneity from conformational ensembles and environmental fluctuations, which broaden signals and complicate interpretation, as seen in single-molecule studies of . Additionally, water's strong mid-infrared absorption interferes with vibrational spectra of hydrated biomolecules, necessitating techniques like photothermal detection to mitigate background noise in aqueous environments.

In Materials Science

Time-resolved spectroscopy has become essential in for probing ultrafast electronic, charge, and in semiconductors, , and photovoltaic materials, enabling insights into processes that govern device performance on to timescales. In perovskites, widely used in solar cells, time-resolved (TRPL) and transient (TA) spectroscopy reveal charge recombination mechanisms, including trap-state-mediated losses that limit , with recombination often spanning hundreds of picoseconds to nanoseconds depending on defect . For instance, studies on halide perovskites highlight how non-radiative recombination via traps reduces carrier , informing strategies to enhance radiative . Similarly, hot carrier cooling in semiconductors like CdSe quantum dots occurs on timescales, as measured by femtosecond TA, where excess energy dissipates via interactions, influencing energy loss in optoelectronic devices. Exciton dynamics in quantum dots, such as CdSe or InP/ZnSe/ZnS variants, are characterized using spectroscopy, which tracks relaxation, recombination, and delocalization processes occurring in picoseconds, with effects modulating lifetimes from tens to hundreds of picoseconds to minimize non-radiative decay. Defect trapping, particularly hole trapping at oxygen vacancies in ZnO nanoparticles, unfolds within 80 picoseconds as probed by time-resolved , altering charge separation and transport in . In thin films, photoinduced phase transitions—such as insulator-to-metal shifts in VO2—are resolved on subpicosecond scales via time-resolved and far-infrared , capturing lattice and electronic rearrangements that enable switchable . Advanced techniques like time-resolved (TR-ARPES) map band structures in , revealing momentum-dependent carrier relaxation and dynamics post-photoexcitation, with cooling times under 100 femtoseconds. Attosecond streaking further dissects conduction band dynamics in materials like , resolving band-gap evolution and carrier motion with ~450-attosecond resolution, highlighting electron-phonon coupling in . These insights directly optimize and LEDs; for example, in cells and LEDs, TR identifies recombination bottlenecks to boost power conversion efficiencies beyond 25% and external quantum efficiencies over 20%, guiding defect passivation and for enhanced stability and performance.

References

  1. [1]
    (PDF) Time‐Resolved Spectroscopy: Instrumentation and Applications
    In this article, we first give an introduction to the fundamentals of time‐resolved electronic absorption, spontaneous excited‐state resonance Raman, and ...
  2. [2]
    Ahmed Zewail – Nobel Lecture - NobelPrize.org
    Ahmed Zewail - Nobel Lecture: Femtochemistry: Atomic-Scale Dynamics of the Chemical Bond Using Ultrafast Lasers · Nobel Prize in Chemistry 1999 · Summary ...
  3. [3]
    Time-resolved Spectroscopy - RP Photonics
    This method is based on the principle of pump–probe measurements, then also called pump–probe spectroscopy. Here, a sample is excited with a pump pulse, ...What is Time-resolved... · Transient Absorption... · Coherent Time-resolved...
  4. [4]
    [PDF] Introduction to time-resolved spectroscopy
    The principle of time-correlated single-photon counting: laser pulse is split into two, one of which starts the time-to-voltage converter, whilst the other ...
  5. [5]
    Time-Resolved Spectroscopy - an overview | ScienceDirect Topics
    Time-resolved spectroscopy is defined as a set of techniques that utilize different observables to probe dynamic processes in molecules over short timescales, ...
  6. [6]
    Flash photolysis and spectroscopy. A new method for the study of ...
    A new technique of flash photolysis and spectroscopy has been developed, using gas-filled flash discharge tubes of very high power.
  7. [7]
    Sub-picosecond spectroscopy - American Institute of Physics
    1970. YEAR. 1975. Decrease in laser-pulse duration during the last decade. In ... E. P. Ippen, C. V. Shank, A. Dienes, Appl. Phys. Lett. 21,348(1972). 5 ...
  8. [8]
    Laser Femtochemistry - Science
    Abstract. Femtochemistry is concerned with the very act of the molecular motion that brings about chemistry, chemical bond breaking, or bond formation on the ...
  9. [9]
    Back to Basics: Ultrafast lasers - Laser Focus World
    The magic is gone. The commercial introduction of modelocked Ti:sapphire lasers in 1990 by Spectra-Physics (Mountain View, CA) began a revolution in the ...
  10. [10]
    Attosecond metrology - Nature
    Nov 29, 2001 · Here we trace electronic dynamics with a time resolution of ≤ 150 as by using a subfemtosecond soft-X-ray pulse and a few-cycle visible light pulse.
  11. [11]
    Attosecond science based on high harmonic generation from gases ...
    Jun 2, 2020 · In this Perspective, we discuss the recent advancements and future trends in generating and characterizing soft X-ray pulses from gas-phase HHG and extreme ...
  12. [12]
    Ultrafast Laser Pulse Generation by Mode Locking - ACS Publications
    Jan 24, 2023 · A technique that enables a fixed phase relationship between laser modes, resulting in laser pulses with very short duration (in the fs or ps range).
  13. [13]
    Advanced time-resolved absorption spectroscopy with an ultrashort ...
    May 11, 2021 · Abstract. Ultrashort visible-near infrared (NIR) pulse generation and its applications to ultrafast spectroscopy are discussed.Missing: I_probe | Show results with:I_probe
  14. [14]
    Compression of amplified chirped optical pulses - ScienceDirect.com
    We have demonstrated the amplification and subsequent recompression of optical chirped pulses. A system which produces 1.06 μm laser pulses widths of 2 ps ...Missing: seminal | Show results with:seminal
  15. [15]
  16. [16]
    High-efficiency near-infrared optical parametric amplifier for intense ...
    Sep 29, 2022 · The OPA has a total conversion efficiency of > 55%, which is the highest value reported to date, with an excellent energy-stability of 0.7% RMS ...
  17. [17]
    Supercontinuum generation in bulk solid-state material with bursts of ...
    Mar 25, 2024 · The presented results identify the fundamental and practical issues for supercontinuum generation in solid-state materials using femtosecond ...
  18. [18]
    Pump–probe Measurements - RP Photonics
    Pump–probe measurements can be used to obtain information on ultrafast phenomena, typically on picosecond or femtosecond time scales.What are Pump–probe... · Characterization of Saturable...
  19. [19]
    Time-Resolved Diffuse Optical Spectroscopy and Imaging Using ...
    In this article, we review time-resolved diffuse optical imaging (TR-DOI) systems using solid-state detectors with a special focus on Single-Photon Avalanche ...
  20. [20]
    Response of a lock-in amplifier to noise | American Journal of Physics
    Aug 1, 2014 · Our results show how a lock-in amplifier can be used to measure the spectral density of noise in the case of a noise-only input signal. Some ...
  21. [21]
    Time-resolved luminescence spectroscopy by the optical Kerr-gate ...
    Oct 15, 2000 · An advantage of the OKG method relative to the conventional up-conversion technique is its ability to directly obtain time-resolved luminescence ...Missing: delay lines ASOPS
  22. [22]
    Ultrafast Fluorescence Spectroscopy via Upconversion - NIH
    In this technique, the fluorescence excited by an ultrafast laser pulse is mixed with another (delayed) portion of the laser pulse in a nonlinear optical ...Missing: ASOPS | Show results with:ASOPS
  23. [23]
    High-speed asynchronous optical sampling with sub-50fs time ...
    We conclude that the presented high-speed ASOPS spectroscopy system matches the time resolution of most conventional time-domain spectroscopy setups.
  24. [24]
    Global and target analysis of time-resolved spectra - ScienceDirect
    The IRF limits the fastest response observable in the experiment. With pumpprobe spectroscopy the IRF is given by the convolution of pump and probe pulses.
  25. [25]
    [PDF] An Introduction to Processing, Fitting, and Interpreting Transient ...
    Feb 16, 2024 · Transient absorption (TA) spectroscopy is a time-resolved spectroscopic technique that monitors the evolution of photo- excited species through ...
  26. [26]
  27. [27]
  28. [28]
    Time-resolved fluorescence spectroscopy - PubMed - NIH
    Time-resolved fluorescence spectroscopy is used to monitor molecular interactions and motions that occur in the picosecond-nanosecond time range.
  29. [29]
    Time-Correlated Single Photon Counting - ScienceDirect.com
    Time-correlated single-photon counting, TCSPC, is a well-established technique for the determination of fluorescence lifetimes and related time-resolved ...
  30. [30]
    [PDF] Time-Correlated Single Photon Counting - PicoQuant
    Figure 1: Measurement of start-stop times in time-resolved fluorescence measurement with TCSPC. ... Time-correlated Single Photon Counting”, Academic Press, ...
  31. [31]
    64 picosecond time resolved time-correlated single photon counting ...
    Feb 1, 2024 · Time-correlated single photon counting (TCSPC) is a technique used to record the arrival time of a photon detected by SPDs. TCSPC is based on ...INTRODUCTION · II. METHODS · III. EXPERIMENTAL SETUP · IV. RESULTS
  32. [32]
    [PDF] Time Resolved Fluorescence Spectroscopy
    Mar 29, 2017 · 3) Learn and apply the mathematical procedures called convolution (and “deconvolution”). I. Background. A. Introduction to Time-Resolved Methods.Missing: fundamentals | Show results with:fundamentals
  33. [33]
    Time-Resolved Fluorescence Anisotropy and Molecular Dynamics ...
    Jan 19, 2021 · In fluorescence anisotropy studies, the rotational mobility of the fluorophore is described by the rotational correlation time, which accounts ...
  34. [34]
    Fluorescence Anisotropy and Resonance Energy Transfer
    The method of time-resolved stopped flow FRET is quite promising for correlating structural dynamics with specific steps in a protein conformational or ...Fluorescence Resonance... · Figure 1 · Single Molecule Fret And Fa
  35. [35]
    Time resolved fluorescence and solvatochromism in donor ...
    Both the absorption and fluorescence spectra show a strong solvatochromic shift indicative of an increased dipole moment in the excited state resulting from ...
  36. [36]
    Time-resolved resonance Raman spectroscopy of photobiological ...
    This article is cited by 27 publications. Anne Myers Kelley. Resonance Raman and Resonance Hyper-Raman Intensities: Structure and Dynamics of Molecular Excited ...Missing: seminal | Show results with:seminal
  37. [37]
    Time-resolved resonance Raman spectroscopy of ... - PNAS
    A simple technique is described that uses a continuous wave laser with electromechanical modulation to obtain time-resolved Raman spectra of transient ...Missing: seminal paper
  38. [38]
    Picosecond UV Kerr‐gated Raman spectroscopy with 280 nm laser ...
    Dec 8, 2022 · We demonstrate a picosecond UV Kerr-gated Raman spectroscopy setup to suppress the fluorescence in Raman spectra excited by 280 nm laser pulses.INTRODUCTION · MATERIAL AND METHODS · RESULTS AND DISCUSSION
  39. [39]
    Time-gated Raman spectroscopy – a review - IOPscience
    Oct 27, 2020 · Time-gated Raman spectroscopy (TG RS) uses precise synchronization to collect Raman signals during short laser pulses, rejecting fluorescence. ...
  40. [40]
    Tracking Ultrafast Structural Dynamics by Time-Domain Raman ...
    Jun 7, 2021 · The pump pulse acts as the Raman pump, whereas the probe pulse works as the Raman probe that monitors the temporal modulation of the absorbance ...
  41. [41]
    Excited state structural evolution during charge-transfer reactions in ...
    Jan 4, 2016 · Picosecond time-resolved Raman spectroscopy was used to follow cooling and intramolecular vibrational ... probe the excited state vibrations ...
  42. [42]
    Ultrabroadband Midinfrared Pump-Probe Spectroscopy Using ...
    May 28, 2015 · We demonstrate ultrabroadband and ultrafast midinfrared (MIR) pump-probe spectroscopy by employing the chirped-pulse up-conversion technique.
  43. [43]
    Time-Resolved Step-Scan FT-IR Spectroscopy - ACS Publications
    Step-scan FT-IR 2 is a time-resolved spectroscopic technique combining laser excitation with fast IR differential detection. Since it detects small changes in ...Introduction · Experimental Section · Conclusion · Author Information
  44. [44]
    Implementation of Time-Resolved Step-Scan Fourier Transform ...
    Time-resolved step-scan Fourier transform infrared (FT-IR) spectroscopy has been shown to be invaluable for studying excited-state structures and dynamics ...
  45. [45]
    Vibrational population relaxation of carbon monoxide in the heme ...
    Vibrational population relaxation times of 335 ± 115 ps for the δ-tautomer and 640 ± 185 ps for the ɛ-tautomer were estimated by using the Landau–Teller model.
  46. [46]
    Energy transfer within the hydrogen bonding network of water ...
    We investigate the dynamic energy flow in the hydrogen bond network of liquid water by a pump-probe experiment.
  47. [47]
    Extracting 2D IR frequency-frequency correlation functions from two ...
    Aug 15, 2011 · The center line slope (CLS) method is often used to extract the frequency-frequency correlation function (FFCF) from 2D IR spectra to delineate
  48. [48]
    Effects of Vibrational Frequency Correlations on Two-Dimensional ...
    Therefore, 2D IR spectroscopy has the capability to reveal the frequency correlations within the same vibrator or between vibrators on different sites. It ...Missing: 2D- spectroscopy
  49. [49]
    Infrared Difference Spectroscopy of Proteins: From Bands to Bonds
    Mar 23, 2020 · After static or time-resolved IR difference spectra are obtained, bands need to be assigned to vibrational modes (section 6). Although bands can ...
  50. [50]
    Dynamics of intramolecular vibrational redistribution in ...
    In spectroscopy, IVR manifests itself in the relaxation of the initially prepared bright state, which is usually a fundamental or an overtone excitation of a ...
  51. [51]
    Antenna-coupled infrared nanospectroscopy of intramolecular ...
    May 8, 2023 · We pursue a qualitatively different approach through femtosecond infrared vibrational nanospectroscopy of molecules coupled to an infrared active antenna.
  52. [52]
  53. [53]
    Time-resolved ARPES studies of quantum materials | Rev. Mod. Phys.
    Feb 27, 2024 · Time-resolved ARPES (TR-ARPES) can probe ultrafast electron dynamics and the out-of-equilibrium electronic structure, providing a wealth of information.
  54. [54]
    Attosecond intra-valence band dynamics and resonant ... - Nature
    Jun 7, 2021 · Time-resolved photoelectron spectroscopy with attosecond precision provides new insights into the photoelectric effect and gives information ...
  55. [55]
    Two‐Photon Photoemission Spectroscopy for Studying Energetics ...
    Apr 3, 2019 · In this review, we will demonstrate the feasibility of time-resolved two-photon photoemission spectroscopy (tr-2PPE) for determination of the ...Missing: seminal | Show results with:seminal
  56. [56]
    Review Theory of two-photon photoemission spectroscopy of surfaces
    We review the various theoretical approaches that have recently been developed to interpret the two-photon photoemission (2PPE) spectra from clean and ...Missing: seminal | Show results with:seminal
  57. [57]
    Femtosecond time-resolved two-photon photoemission studies of ...
    May 8, 2019 · Femtosecond time-resolved two-photon photoemission spectroscopy has been used to probe the electronic structure and dynamics of photoexcited charge carriers.
  58. [58]
    One-step theory of two-photon photoemission | Phys. Rev. B
    The goal of the present paper is to apply the general formalism to the setup of 2PPE experiments. In this case we can benefit from the fact that the initial ...Missing: seminal | Show results with:seminal
  59. [59]
    Decay and dephasing of image-potential states studied by time ...
    In a two-photon photoemission experiment a first laser pulse excites the electron into an intermediate state while a second pulse emits the electron into the ...
  60. [60]
    Analysis of two-photon photoemission from Si(001) | Phys. Rev. B
    Dec 19, 2012 · We have applied our ab initio simulation approach for the photoemission process at solid surfaces to calculate two-photon photoemission ...Missing: seminal | Show results with:seminal
  61. [61]
    Optical decoherence and quantum beats in Cs/Cu(111)
    An unoccupied state due to the s–p or s–d anti-bonding hybridization of the alkali atoms has been identified by inverse and two-photon photoemission techniques ...
  62. [62]
    Ultrafast energizing the parity-forbidden dark exciton in black ... - NIH
    Apr 29, 2025 · Interferometric time- and angle-resolved two-photon photoemission spectroscopy reveals dark exciton dynamics of black phosphorus on ~100 fs ...
  63. [63]
    Four Wave Mixing spectroscopy at the interface between the time ...
    We demonstrate a new approach to Four Wave Mixing spectroscopy involving simultaneous measurements at time and frequency domains, where spectral selectivity is ...
  64. [64]
    The transient grating: a holographic window to dynamic processes
    This article is cited by 65 publications. Daphné Lubert-Perquel, Swagata Acharya, Justin C. Johnson. Optically Addressing Exciton Spin and Pseudospin in ...Missing: Heritage original
  65. [65]
    [PDF] Transient Grating Spectroscopy of Photocarrier Dynamics in ... - arXiv
    In this paper, we use laser-induced transient grating (TG) spectroscopy to characterize thin films of the archetypal organic semiconductor regioregular poly(3- ...
  66. [66]
    Photon-Echo Spectroscopy of Organic Mixed Crystals | Phys. Rev. Lett.
    Jun 7, 1976 · Abstract. We report the first observation of photon echoes in an organic mixed crystal. The echo was observed after excitation with a nitrogen- ...Missing: original paper
  67. [67]
    Photon echo studies of photosynthetic light harvesting - PMC
    Photon echo peak shift spectroscopy can be used to determine the strength of coupling between a pigment and its surrounding environment including neighboring ...Photon Echo Peak Shift... · Two-Dimensional Fourier... · ApplicationsMissing: seminal papers<|separator|>
  68. [68]
    Exciton Analysis in 2D Electronic Spectroscopy - ACS Publications
    Three-pulse photon echo spectroscopy is performed using a phase-matched box geometry and heterodyne detection. We employ diffractive optics to achieve long-term ...
  69. [69]
    Evidence for wavelike energy transfer through quantum coherence ...
    Apr 12, 2007 · In the present experiment, we use two-dimensional electronic spectroscopy to observe oscillations caused by electronic coherence evolving during ...Missing: 2D electronic spectroscopy FMO
  70. [70]
    [PDF] Optical multidimensional coherent SPECTROSCOPY - Shaul Mukamel
    That pulse sequence, with the phase-conjugated pulse arriving first, produces a so-called rephasing spectrum, which is only one of several possible 2D spectra.
  71. [71]
    Coherent radiative control of unimolecular reactions: Selective bond ...
    We show that the principle of coherent control of chemical reactions can be implemented, in a two pulse sequence, to control the relative yield of reaction ...<|separator|>
  72. [72]