<<

Unit-2 Fourier Transform

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Unit-2 Fourier Transform Spectroscopy

Measuring an emission : Fourier transform spectroscopy

Unit-2 Fourier Transform Spectroscopy

An example of a spectrum: The spectrum of emitted by the blue of a butane torch. The horizontal axis is the of light, and the vertical axis represents how much light is emitted by the torch at that wavelength. One of the most basic tasks in spectroscopy is to characterize the spectrum of a light source: how much light is emitted at each different wavelength. The most straightforward way to measure a spectrum is to pass the light through a , an instrument that blocks all of the light except the light at a certain wavelength (the un-blocked wavelength is set by a knob on the monochromator). Then the intensity of this remaining (single-wavelength) light is measured. The measured intensity directly indicates how much light is emitted at that wavelength. By varying the monochromator's wavelength setting, the full spectrum can be measured. This simple scheme in fact describes how some work. Fourier-transform spectroscopy is a less intuitive way to get the same information. Rather than allowing only one wavelength at a time to pass through to the detector, this technique lets through a beam containing many different of light at once, and measures the total beam intensity. Next, the beam is modified to contain a different combination of wavelengths, giving a second data point. This process is repeated many times. Afterwards, a computer takes all this data and works backwards to infer how much light there is at each wavelength. To be more specific, between the light source and the detector, there is a certain configuration of mirrors that allows some wavelengths to pass through but blocks others (due to wave interference). The beam is modified for each new data point by moving one of the mirrors; this changes the set of wavelengths that can pass through.

As mentioned, computer processing is required to turn the raw data (light intensity for each mirror position) into the desired result (light intensity for each wavelength). The processing required turns out to be a common algorithm called the Fourier transform (hence the name, "Fourier-transform spectroscopy"). The raw data is sometimes called an "interferogram". Because of the existing computer equipment requirements, and the ability of light to analyze very small amounts of substance, it is often beneficial to automate many aspects of the sample preparation. The sample can be better preserved and the results are much easier to replicate. Both of these benefits are important, for instance, in testing situations that may later involve legal action, such as those involving drug specimens.[2] Unit-2 Fourier Transform Spectroscopy

Measuring an absorption spectrum

An "interferogram" from a Fourier-transform . This is the "raw data" which can be Fourier-transformed into an actual spectrum. The peak at the center is the ZPD position ("zero path difference"): Here, all the light passes through the interferometer because its two arms have equal length.

The method of Fourier-transform spectroscopy can also be used for . The primary example is "FTIR Spectroscopy", a common technique in chemistry. In general, the goal of absorption spectroscopy is to measure how well a sample absorbs or transmits light at each different wavelength. Although absorption spectroscopy and emission spectroscopy are different in principle, they are closely related in practice; any technique for emission spectroscopy can also be used for absorption spectroscopy. First, the of a broadband lamp is measured (this is called the "background spectrum"). Second, the emission spectrum of the same lamp shining through the sample is measured (this is called the "sample spectrum"). The sample will absorb some of the light, causing the spectra to be different. The ratio of the "sample spectrum" to the "background spectrum" is directly related to the sample's absorption spectrum. Accordingly, the technique of "Fourier-transform spectroscopy" can be used both for measuring emission spectra (for example, the emission spectrum of a ), and absorption spectra (for example, the absorption spectrum of a liquid).

Applications of Fourier transform spectroscopy Unit-2 Fourier Transform Spectroscopy

Forensic chemists rely on a multitude of instruments to identify unknown substances found at a scene.[26] Different methods can be used to determine the identity of the same substance, and it is up to the examiner to determine which method will produce the best results. Factors that forensic chemists might consider when performing an examination are the length of time a specific instrument will take to examine a substance and the destructive nature of that instrument. They prefer using nondestructive methods first, to preserve the evidence for further examination.[27] Nondestructive techniques can also be used to narrow down the possibilities, making it more likely that the correct method will be used the first time when a destructive method is used.[27]

Spectroscopy

The two main standalone spectroscopy techniques for forensic chemistry are FTIR and AA spectroscopy. FTIR is a nondestructive process that uses light to identify a substance. The attenuated total reflectance sampling technique eliminates the need for substances to be prepared before analysis.[28] The combination of nondestructiveness and zero preparation makes ATR FTIR analysis a quick and easy first step in the analysis of unknown substances. To facilitate the positive identification of the substance, FTIR instruments are loaded with databases that can be searched for known spectra that match the unknown's spectra. FTIR analysis of mixtures, while not impossible, presents specific difficulties due to the cumulative nature of the response. When analyzing an unknown that contains more than one substance, the resulting spectra will be a combination of the individual spectra of each component.[29] While common mixtures have known spectra on file, novel mixtures can be difficult to resolve, making FTIR an unacceptable means of identification. However, the instrument can be used to determine the general chemical structures present, allowing forensic chemists to determine the best method for analysis with other instruments. For example, a methoxy group will result in a peak between 3,030 and 2,950 wavenumbers (cm−1).[30]

Atomic absorption spectroscopy (AAS) is a destructive technique that is able to determine the elements that make up the analyzed sample. AAS performs this analysis by subjecting the sample to an extremely high heat source, breaking the atomic bonds of the substance, leaving free . Radiation in the form of light is then passed through the sample forcing the atoms to jump to a higher state.[31]:2 Forensic chemists can test for each element by using a Unit-2 Fourier Transform Spectroscopy corresponding wavelength of light that forces that element's atoms to a higher energy state during the analysis.[31]:256 For this reason, and due to the destructive nature of this method, AAS is generally used as a confirmatory technique after preliminary tests have indicated the presence of a specific element in the sample. The concentration of the element in the sample is proportional to the amount of light absorbed when compared to a blank sample.[32] AAS is useful in cases of suspected heavy metal poisoning such as with arsenic, lead, mercury, and cadmium. The concentration of the substance in the sample can indicate whether heavy metals were the cause of death.[33]

NMR spectroscopy

NMR spectroscopy is one of the principal techniques used to obtain physical, chemical, electronic and structural information about due to the chemical shift of the resonance of the nuclear spins in the sample. Peak splittings due to J- or dipolar couplings between nuclei are also useful. NMR spectroscopy can provide detailed and quantitative information on the functional groups, topology, dynamics and three-dimensional structure of molecules in solution and the state. Since the area under an NMR peak is usually proportional to the number of spins involved, peak integrals can be used to determine composition quantitatively. Structure and molecular dynamics can be studied (with or without "magic angle" spinning (MAS)) by NMR of quadrupolar nuclei (that is, with S > 1/2) even in the presence of magnetic "dipole-dipole" interaction broadening (or simply, dipolar broadening) which is always much smaller than the quadrupolar interaction strength because it is a magnetic vs. an electric interaction effect. Additional structural and chemical information may be obtained by performing double-quantum NMR experiments for pairs of spins or quadrupolar nuclei such as. Furthermore, nuclear magnetic resonance is one of the techniques that has been used to design quantum automata, and also build elementary quantum computers.[11][12]

Continuous-wave (CW) spectroscopy

In the first few decades of nuclear magnetic resonance, spectrometers used a technique known as continuous-wave (CW) spectroscopy, where the transverse spin magnetization generated by a weak oscillating magnetic field is recorded as a function of the oscillation or static Unit-2 Fourier Transform Spectroscopy

[9] field strength B0. When the oscillation frequency matches the nuclear resonance frequency, the transverse magnetization is maximized and a peak is observed in the spectrum. Although NMR spectra could be, and have been, obtained using a fixed constant magnetic field and sweeping the frequency of the oscillating magnetic field, it was more convenient to use a fixed frequency source and vary the current (and hence magnetic field) in an electromagnet to observe the resonant absorption signals. This is the origin of the counterintuitive, but still common, "high field" and "low field" terminology for low frequency and high frequency regions, respectively, of the NMR spectrum.

As of 1996, CW instruments were still used for routine work because the older instruments were cheaper to maintain and operate, often operating at 60 MHz with correspondingly weaker (non- superconducting) electromagnets cooled with water rather than liquid . One radio coil operated continuously, sweeping through a range of frequencies, while another orthogonal coil, designed not to receive radiation from the transmitter, received signals from nuclei that reoriented in solution. As of 2014, low-end refurbished 60 MHz and 90 MHz systems were sold as FT-NMR instruments, and in 2010 the "average workhorse" NMR instrument was configured for 300 MHz.

CW spectroscopy is inefficient in comparison with Fourier analysis techniques (see below) since it probes the NMR response at individual frequencies or field strengths in succession. Since the NMR signal is intrinsically weak, the observed spectrum suffers from a poor signal-to-noise ratio. This can be mitigated by signal averaging, i.e. adding the spectra from repeated measurements. While the NMR signal is the same in each scan and so adds linearly, the random noise adds more slowly – proportional to the square root of the number of spectra (see random walk). Hence the overall signal-to-noise ratio increases as the square-root of the number of spectra measured.

Fourier-transform spectroscopy

Most applications of NMR involve full NMR spectra, that is, the intensity of the NMR signal as a function of frequency. Early attempts to acquire the NMR spectrum more efficiently than simple CW methods involved illuminating the target simultaneously with more than one Unit-2 Fourier Transform Spectroscopy frequency. A revolution in NMR occurred when short radio-frequency pulses began to be used, with a frequency centered at the middle of the NMR spectrum. In simple terms, a short pulse of a given "carrier" frequency "contains" a range of frequencies centered about the carrier frequency, with the range of excitation (bandwidth) being inversely proportional to the pulse duration, i.e. the Fourier transform of a short pulse contains contributions from all the frequencies in the neighborhood of the principal frequency. The restricted range of the NMR frequencies made it relatively easy to use short (1 - 100 microsecond) radio frequency pulses to excite the entire NMR spectrum.

Applying such a pulse to a set of nuclear spins simultaneously excites all the single-quantum NMR transitions. In terms of the net magnetization vector, this corresponds to tilting the magnetization vector away from its equilibrium position (aligned along the external magnetic field). The out-of-equilibrium magnetization vector then precesses about the external magnetic field vector at the NMR frequency of the spins. This oscillating magnetization vector induces a voltage in a nearby pickup coil, creating an electrical signal oscillating at the NMR frequency. This signal is known as the free induction decay (FID), and it contains the sum of the NMR responses from all the excited spins. In order to obtain the frequency-domain NMR spectrum (NMR absorption intensity vs. NMR frequency) this time-domain signal (intensity vs. time) must be Fourier transformed. Fortunately, the development of Fourier transform (FT) NMR coincided with the development of digital computers and the digital Fast Fourier Transform. Fourier methods can be applied to many types of spectroscopy.

Richard R. Ernst was one of the pioneers of pulsed NMR and won a Nobel Prize in chemistry in 1991 for his work on Fourier Transform NMR and his development of multi-dimensional NMR spectroscopy