Spectroscopy for Materials Characterization. Группа авторов

Чтение книги онлайн.

Читать онлайн книгу Spectroscopy for Materials Characterization - Группа авторов страница 49

Spectroscopy for Materials Characterization - Группа авторов

Скачать книгу

from the femtosecond laser. The first pulse hitting the sample is the pump, which is resonant with a certain transition of the sample and photoexcites it. To distinguish this pulse from the following two, it is often named actinic or photochemical pump. After that, a combination of two more pulses is used to probe the vibrational mode pattern of the excited system. One acts as Raman pump pulse, long in the time domain, hence narrowband in frequency around ω, as required for spectral resolution; the third pulse, called the Raman probe beam, provides the seed responsible of stimulating the Raman emission. The two pulses are overlapping in time and hit the sample together after a certain delay t from the pump. However, the Raman seed is much shorter in time (<100 fs), hence spectrally broadband and contains nonzero spectral density across a wide interval of ω . In these conditions, one observes a strong enhancement of the Raman scattering at ω , which can amount to several orders of magnitude. If the Raman probe spectrum is dispersed and measured with and without the Raman pump, one can directly obtain the stimulated Raman spectrum as the gain in the probe, that is the natural logarithm of the ratio: (Probe)Pump on/(Probe)Pump off. Besides, the detection is greatly facilitated by the fact that the stimulated Raman signal is emitted in a defined direction. In fact, femtosecond stimulated Raman spectroscopy allows to acquire high‐quality Raman spectra even in the presence of a strong fluorescence [52, 61].

Image described by caption

      Source: Reprinted with permission from [55]. Copyright (1999) by the American Physical Society.

      Panel (b): Energy level diagram for time‐resolved stimulated Raman.

      Source: Adapted with permission from McCamant et al. [60]. Copyright (2003) American Chemical Society.

      Panel (c): One of the earliest time‐resolved stimulated experiment, probing the trans‐cis photoisomerization dynamics of 4‐(dicyanomethylene)‐2‐methyl‐6‐(4‐dimethylaminostyryl)‐4H‐pyran, dissolved in a 3 mM concentration in DMSO and photoexcited at 397 nm. The Raman pump pulse was tuned at 795 nm (bandwidth 1.5 nm), while the Raman probe pulse, obtained by supercontinuum generation, covered the region 600–1000 nm. Arrows indicate the Raman signals appearing due to the photogenerated transient states, while asterisks indicate the Raman lines of the solvent.

      Source: Reprinted figure with permission from Yoshizawa and Kurosawa [55]. Copyright (1999) by the American Physical Society.

      What is more important, however, is that the use of this detection approach provides the trick to achieve very high temporal resolution [61]. In fact, despite Raman scattering occurring during the whole duration of the Raman pump pulse (∼1 ps or more), it is only within the duration of the short‐lived Raman probe beam that stimulated Raman process occurs. Therefore, the time resolution of the measurement is now controlled by the femtosecond duration of the probe pulse. For example, a narrowband Raman pump pulse at 560 nm, with duration of several picoseconds, can be combined with a probe pulse centered, say, at 580 nm with a FWHM of 20 nm, which is broad enough to stimulate Raman emission over a relatively wide spectral range (∼1000 cm−1), and can be compressed down to <25 fs.

      While femtosecond stimulated Raman can be used as a way to record a ground state Raman spectrum, the final goal of these experiments is usually following the evolution initiated by the actinic pump, with time resolution given by the cross‐correlation of the actinic and Raman probe pulses, which can reach a few tens of femtoseconds. To do so, the stimulated Raman spectrum for the unexcited sample is compared to that obtained at variable delays from photoexcitation, in order to follow the effect of relaxation. The result of the experiment can be plotted either in terms of an absolute, time‐dependent Raman spectrum (e.g. as in Figure 3.6c), which contains both the features of the unexcited ground molecules and those of the excited molecules, or as a difference Raman spectrum, calculated with respect to the unexcited sample.

Schematic illustration of a possible configuration for a time-resolved stimulated Raman experiment.

      Source: Adapted with permission from McCamant et al. [60]. Copyright (2003) American Chemical Society.

      Several variants of this general scheme are possible. In some setups, for example, tunable Raman pump and probe pulses are both obtained by NOPAs [62]. In other cases, the fundamental beam from the amplifier was directly used as the Raman probe [63]. In regard to detection, some setups make use of a reference beam [60]: the probe is split in two in order to generate a reference beam, allowing for efficient shot‐to‐shot normalization of white light fluctuations. Experiments have been conducted over a wide range of Raman pump wavelengths, from the near IR to the UV [64]. This wide flexibility also allows to tune the Raman pump for pre‐resonant [63] and resonant [53] Raman process, which provides a convenient route to further enhance the Raman signal, and to single out the Raman contribution from the chromophore of interest.

      In regard to data acquisition, several pulse chopping sequences can be used in this type of experiments [52]. By a simple chopper in the Raman pump path, one can acquire the stimulated Raman spectrum from the ratio between Raman‐pumped and Raman‐unpumped spectra cyclically recorded by the array detector. Then, the Raman spectra of the ground state and photoexcited sample can be obtained sequentially, and compared to each other, by

Скачать книгу