You are on page 1of 15

Infrared Spectroscopy

Review: High Energy = High Frequency = Short Wavelength E = h = hc/ Wavenumber = reciprocal centimeters (cm-1) Range for vibrational IR radiation = 2.5-25 m wavelengths 1 cm-1 = (1/m) 10000 1 m = (1/ cm-1) 10000 Calculate the range of wavenumbers corresponding an IR spectrum! Region of Spectrum X-Ray UV/Visible Infrared Microwave Radiofrequency Why IR? Wavelength < 200 nm 200-800 nm 2.5-25 m > 25 m MW < 1 m 1-5 m Transition Bond breaking Electronic Vibrational Rotational NMR/ESR

Molecules absorb energy corresponding to the natural vibrational frequencies of bond stretching and bending. Absorption: When the IR frequency = natural vibrational frequency Result: Amplitude of signal increases! Requirement: Only bonds with a dipole moment may absorb energy. Which of the following bonds would you expect to absorb energy in the IR? Whats observed?
H O H Cl Cl O C O H H

Br

Cl

C H

CH3 H

Stretching: Symmetric (High E) Asymmetric (Lower E) Bending: (Real low E)

= frequency in cm -1 = 1 2c k k = force constant m1m2 = m1 + m2 = reduced mass k

Stronger bonds = Higher frequency = Higher wavenumber C 2150 C C C C C 1650 1200 (cm-1)

Bigger atoms = Bigger reduced mass = Lower wavenumber C H 3000 C C C O C Cl C Br 600 (cm-1)

1200

1100

750

Scissoring Wagging Rocking Twisting Fundamental Absorptions = Excitation from the ground state to the first excited state. Overtones = Excitation from the ground state to higher energy states.

Utility:
Bond

Wavenumber O-H N-H C-H C N 3400 3400 3000 2250

Bond C C C=O C=C C-O

Wavenumber 2150 1710 1650 1100

Natural Frequency of Vibration of a Bond: Calculating approximate IR absorptions:


= 1 2c k = 4.12 k 1 dyne = 1.020 X 10-3 g M1 + M 2 M1 and M2 = atomic weights = M1M2

k = 5 X 105 dynes/cm k = 10 X 105 dynes/cm k = 15 X 105 dynes/cm

carbon single bond double bond triple bond

Calculate the theoretical absorption frequency for a carbonyl group.

INFRARED SPECTROMETER = SPECTROPHOTOMETER 1.) Dispersive: a.) b.) c.) d.) Double-Beam: The IR radiation from a source is split into two beams. One of the beams passes through the sample while the other beam passes through a reference. Selector: Both beams (reference and sample) are passed to the diffraction grating as a combined beam. Diffraction Grating: Individual wavelengths of the combined IR beam are passed to the detector one at a time. Detector: Measures the difference between the sample versus the reference portion of the combined beam to yield just the sample contribution.

2.)

Fourier Transform (FT): a.) Single Beam: The IR radiation from the source is sent to a stationary mirror and a moving mirror. The moving mirror accomplishes the same as a reference cell. It creates an interference beam to the sample beam that subtracts out the background. Laser Beam Splitter: The laser produces red light of a known frequency at a known moving mirror position. The frequency enters into the FT for the final spectrum. Interferogram: A plot of intensity versus time. A complex signal with a wavelike pattern that contains all of the frequencies between 4000 cm-1 and 400 cm-1. The FID if you will! Fourier Transform: There is a computer program that takes the wavelike pattern of the interferogram and separates the individual absorption frequencies into a spectrum. FT-IR Only one moving part Entire IR region is detected at once Spectrum = Many scans High s/N

b.) c.) d.)

Dispersed

Many moving parts Scans total IR region 1 by 1 Spectrum = 1 scan Low s/N
Samples

Glass and plastics absorb in the IR range making it necessary to use some other material for sample preparation.

1.

Cells (Salt plates: NaCl or KBr) a.) b.) NaCl: The range for detection is 4000 cm-1 to 650 cm-1. These cells are relatively inexpensive. KBr: These cells cover a wider range of frequencies (4000 cm-1 to 400 cm-1). The down side is the expense.

2. 3.

Liquids: Only one drop of sample is necessary between two salt plates. The problem is that the salt plates are water soluble. Solids: a.) KBr Pellet: A solid sample may be mixed with KBr powder and pressed into a clear pellet for analysis. b.) Nujol Mull: The solid sample may be mixed with mineral oil and placed on a salt plate. Nujol absorbs at 2924 cm-1, 1462 cm-1, and 1377 cm-1. c.) CCl4: The solid may be dissolved in an organic solvent for analysis. Carbon tetrachloride absorbs at 785 cm-1.

To Focus on the Main Points Introduction to IR Spectra Source: www.chem.ucla.edu Theory An invaluable tool in organic structure determination and verification involves the class of electromagnetic (EM) radiation with frequencies between 4000 and 400 cm-1 (wavenumbers). The category of EM radiation is termed infrared (IR) radiation, and its application to organic chemistry known as IR spectroscopy. Radiation in this region can be utilized in organic structure determination by making use of the fact that it is absorbed by interatomic bonds in organic compounds. Chemical bonds in different environments will absorb varying intensities and at varying frequencies. Thus IR spectroscopy involves collecting absorption information and analyzing it in the form of a spectrum. The frequencies at which there are absorptions of IR radiation ("peaks" or "signals") can be correlated directly to bonds within the compound in question. Because each interatomic bond may vibrate in several different motions (stretching or bending), individual bonds may absorb at more than one IR frequency.

Stretching absorptions usually produce stronger peaks than bending, however the weaker bending absorptions can be useful in differentiating similar types of bonds (e.g. aromatic substitution). It is also important to note that symmetrical vibrations do not cause absorption of IR radiation. For example, neither of the carbon-carbon bonds in ethene or ethyne absorb IR radiation. Regions of the IR Spectrum Over time organic chemists have recorded and catalogued the types and locations of IR absorptions produced by a wide variety of chemical bonds in various chemical environments. These data can be quickly referenced through tables of IR absorption ranges and compared to the spectrum under consideration. As a general rule, the most important factors determining where a chemical bond will absorb are the bond order and the types of atoms joined by the bond. Conjugation and nearby atoms shift the frequency to a lesser degree. Therefore the same or similar functional groups in different molecules will typically absorb within the same, specific frequency ranges. Consequently tables of IR absorptions are arranged by functional group -- in some versions these may be further subdivided to give more precise information. In IR absorption tables, signal intensities (height) are usually denoted by the following abbreviations: w = weak, m = medium, s = strong, v = variable. A broad signal shape is sometimes indicated by br. Occasionally absorption frequency is given as a single approximation denoted with an ~ rather than a range. A Table of IR Absorptions These trends in aborption can be further summarized into the following categories Wavenumber Vibration 3600 - 2700 cm-1 X-H 2700 - 1900 cm-1 X=Y 1900 - 1500 cm-1 X=Y 1500 - 500 cm-1 X-Y Upon first inspection, a typical infrared spectrum can be visually divided into two regions.

The left half, above 2000 cm-1, usually contains relatively few peaks, but some very diagnostic information can be found here. First, alkane C-H stretching absorptions just below 3000 cm-1 demonstrate the presence of saturated carbons, and signals just above 3000 cm-1 demonstrate unsaturation. A very broad peak in the region between 3100 and 3600 cm-1 indicates the presence of exchangeable protons, typically from alcohol, amine, amide or carboxylic acid groups (see further discussion of this below). The frequencies from 2800 to 2000 cm-1 are normally void of other absorptions, so the presence of alkyne or nitrile groups can be easily seen here. In contrast, the right half of the spectrum, below 2000 cm-1, normally contains many peaks of varying intensities, many of which are not readily identifiable. Two signals which can be seen clearly in this area is the carbonyl group, which is a very strong peak around 1700 cm-1, and the C-O bond with can be one or two strong peaks around 1200 cm-1. This complex lower region is also known as the "fingerprint region" because almost every organic compound produces a unique pattern in this area -Therefore identity can often be confirmed by comparison of this region to a known spectrum.

Additional IR Concepts Although the above and similar IR absorption tables provide a good starting point for assigning simple IR spectra, it is often necessary to understand in greater detail some more specific properties of IR spectra. The following topics cover some of the most important of these principles. The Effects of Mass on Frequency As mentioned previously, one of the major factors influencing the IR absorption frequency of a bond are the identity of the two atoms involved. To be more precise, it is the masses of the two atoms which are of greater importance. The greater the masses of attached atoms, the lower the IR frequency at which the bond will absorb.

An example of this are the spectra of chloroform and deuterochloroform -notice that the two major differences in these spectra are (1) the disappearance of the C-H stretching (3020 cm-1) and bending (1220 cm-1) in deuterated compound and (2) a shift to the right about 20 cm-1 relative to the CHCl3. The first is caused simply by the lack of C-H bonds in CDCl3. The second is illustrative of this property that heavier atoms (deuterium vs. hydrogen) will cause attached bonds to absorb at lower frequencies. Free vs. Hydrogen-Bonded Hydroxyl Groups One of the most distinct and easily recognizable peaks in an IR spectrum is the broad O-H absorption of alcohols and phenols. However it important to understand why this broadening takes place and to consider the situations in which the peak may not have this characteristic shape. First, note that any significant quantity of a compound will contain a very large number of individual molecules, and each of these may be hydrogen bonded to a slightly different extent. Thus as an IR spectrum is acquired IR absorptions will occur at varying frequencies for each of these bonds. The end result is that the IR peak appears broadened, as it is an average of all these slightly different absorptions. It is possible to acquire IR spectra of hydroxyl-containing compounds without seeing this broad signal. By creating a very dilute solution of the sample, or acquiring the IR spectra in the gas phase, hydrogen bonding is prevented through lack of molecular contact. Even in concentrated solution, larger compounds may sterically hinder hydrogen bonding, preventing exchange. Inthese situations the broad O-H peak is replaced by a sharp signal around 3600 cm-1. This effect can be seen in the IR spectra of t-butanol, dilute and concentrated.

Spectroscopy
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Extremely high resolution spectrograph of the Sun showing thousands of elemental absorption lines (fraunhofer lines) Spectroscopy is the study of the interaction between radiation (electromagnetic radiation, or light, as well as particle radiation) and matter. Spectrometry is the measurement of these interactions and a machine which performs such measurements is a spectrometer or spectrograph. A plot of the interaction is referred to as a spectrogram, or, informally, a spectrum. Historically, spectroscopy referred to a branch of science in which visible light was used for the theoretical study of the structure of matter and for qualitative and quantitative analyses. Recently, however, the definition has broadened as new techniques have been developed that utilise not only visible light, but many other forms of radiation. Spectroscopy is often used in physical and analytical chemistry for the identification of substances through the spectrum emitted from or absorbed by them. Spectroscopy is also heavily used in astronomy and remote sensing. Most large telescopes have spectrometers, which are used either to measure the chemical composition and physical properties of astronomical objects or to measure their velocities from the Doppler shift of their spectral lines.

Infrared spectroscopy
From Wikipedia, the free encyclopedia

(Redirected from IR spectroscopy) Jump to: navigation, search Infrared spectroscopy (IR Spectroscopy) is the subset of spectroscopy that deals with the IR region of the EM spectrum. It covers a range of techniques, with the most common

type by far being a form of absorption spectroscopy. As with all spectroscopic techniques, it can be used to identify a compound and to investigate the composition of a sample. Infrared spectroscopy correlation tables are tabulated in the literature.

Contents
[hide]

1 Theory 2 Sample preparation 3 Typical method 4 Summary of absorptions of bonds in organic molecules 5 Uses and applications 6 Fourier transform infrared spectroscopy 7 Two-dimensional infrared spectroscopy 8 See also 9 References 10 External links

[edit] Theory
The infrared portion of the electromagnetic spectrum is divided into three regions; the near-, mid- and far- infrared, named for their relation to the visible spectrum. The farinfrared, (approx. 400-10 cm-1) lying adjacent to the microwave region, has low energy and may be used for rotational spectroscopy. The mid- infrared (approx. 4000-400 cm-1) may be used to study the fundamental vibrations and associated rotational-vibrational structure, whilst the higher energy near-IR (14000-4000 cm-1) can excite overtone or harmonic vibrations. Infrared spectroscopy works because chemical bonds have specific frequencies at which they vibrate corresponding to energy levels. The resonant frequencies or vibrational frequencies are determined by the shape of the molecular potential energy surfaces, the masses of the atoms and, eventually by the associated vibronic coupling. In order for a vibrational mode in a molecule to be IR active, it must be associated with changes in the permanent dipole. In particular, in the Born-Oppenheimer and harmonic approximations, i.e. when the molecular Hamiltonian corresponding to the electronic ground state can be approximated by a harmonic oscillator in the neighborhood of the equilibrium molecular geometry, the resonant frequencies are determined by the normal modes corresponding to the molecular electronic ground state potential energy surface. Nevertheless, the resonant frequencies can be in a first approach related to the strength of the bond, and the mass of the atoms at either end of it. Thus, the frequency of the vibrations can be associated with a particular bond type.

Simple diatomic molecules have only one bond, which may stretch. More complex molecules may have many bonds, and vibrations can be conjugated, leading to infrared absorptions at characteristic frequencies that may be related to chemical groups. The atoms in a CH2 group, commonly found in organic compounds can vibrate in six different ways, symmetrical and asymmetrical stretching, scissoring, rocking, wagging and twisting; as shown below: Symmetrical stretching Asymmetrical stretching

Scissoring

Rocking

Wagging

In order to measure a sample, a beam of infrared light is passed through the sample, and the amount of energy absorbed at each wavelength is recorded. This may be done by scanning through the spectrum with a monochromatic beam, which changes in wavelength over time, or by using a Fourier transform instrument to measure all wavelengths at once. From this, a transmittance or absorbance spectrum may be plotted, which shows at which wavelengths the sample absorbs the IR, and allows an interpretation of which bonds are present. This technique works almost exclusively on covalent bonds. Clear spectra are obtained from samples with few IR active bonds and high levels of purity. More complex molecular structures lead to more absorption bands and more complex spectra. The technique has been used for the characterization of very complex mixtures however.

[edit] Sample preparation


Gaseous samples require little preparation beyond purification, but a sample cell with a long pathlength (typically 5-10 cm) is used as gases show relatively weak absorbances. Liquid samples can be sandwiched between two plates of a high purity salt (commonly sodium chloride, or common salt, although a number of other salts such as potassium bromide or calcium fluoride are also used). The plates are transparent to the infrared light and will not introduce any lines onto the spectra. Some salt plates are highly soluble in water, and so the sample, washing reagents and the like must be anhydrous (without water).

Solid samples can be prepared in two major ways. The first is to crush the sample with a mulling agent (usually nujol) in a marble or agate mortar, with a pestle. A thin film of the mull is applied onto salt plates and measured. The second method is to grind a quantity of the sample with a specially purified salt (usually potassium bromide) finely (to remove scattering effects from large crystals). This powder mixture is then crushed in a mechanical die press to form a translucent pellet through which the beam of the spectrometer can pass. It is important to note that spectra obtained from different sample preparation methods will look slightly different from each other due to the different physical states the sample is in.

[edit] Typical method

Typical apparatus A beam of infra-red light is produced and split into two separate beams. One is passed through the sample, the other passed through a reference which is often the substance the sample is dissolved in. The beams are both reflected back towards a detector, however first they pass through a splitter which quickly alternates which of the two beams enters the detector. The two signals are then compared and a printout is obtained. A reference is used for two reasons:

This prevents fluctuations in the output of the source affecting the data This allows the effects of the solvent to be cancelled out (the reference is usually a pure form of the solvent the sample is in)

[edit] Summary of absorptions of bonds in organic molecules


Main article: Infrared Spectroscopy Correlation Table

Wavenumbers listed in cm-1.

[edit] Uses and applications


Infrared spectroscopy is widely used in both research and industry as a simple and reliable technique for measurement, quality control, and dynamic measurement. The instruments are now small, and can be transported, even for use in field trials. With increasing technology in computer filtering and manipulation of the results, samples in solution can now be measured accurately (water produces a broad absorbance across the range of interest, and thus renders the spectra unreadable without this computer treatment). Some machines will also automatically tell you what substance is being measured from a store of thousands of reference spectra held in storage. By measuring at a specific frequency over time, changes in the character or quantity of a particular bond can be measured. This is especially useful in measuring the degree of polymerization in polymer manufacture. Modern research machines can take infrared measurements across the whole range of interest as frequently as 32 times a second. This can be done whilst simultaneous measurements are made using other techniques. This makes the observations of chemical reactions and processes quicker and more accurate. Techniques have been developed to assess the quality of tea-leaves using infrared spectroscopy. This will mean that highly trained experts (also called 'noses') can be used more sparingly, at a significant cost saving. Infrared spectroscopy has been highly successfully for applications in both organic and inorganic chemistry. Infrared spectroscopy has also been successfully utilized in the field of semiconductor microelectronics [1]: for example, infrared spectroscopy can be applied to semiconductors like silicon, gallium arsenide, gallium nitride , zinc selenide, amorphous silicon, silicon nitride, etc.

[edit] Fourier transform infrared spectroscopy


Main article: Fourier transform spectroscopy Fourier transform infrared (FTIR) spectroscopy is a measurement technique for collecting infrared spectra. Instead of recording the amount of energy absorbed when the

frequency of the infra-red light is varied (monochromator), the IR light is guided through an interferometer. After passing the sample the measured signal is the interferogram. Performing a mathematical Fourier Transform on this signal results in a spectrum identical to that from conventional (dispersive) infrared spectroscopy. FTIR spectrometers are cheaper than conventional spectrometers because building of interferometers is easier than the fabrication of a monochromator. In addition, measurement of a single spectrum is faster for the FTIR technique because the information at all frequencies is collected simultaneously. This allows multiple samples to be collected and averaged together resulting in an improvement in sensitivity. Because of its various advantages, virtually all modern infrared spectrometers are FTIR instruments.

[edit] Two-dimensional infrared spectroscopy


Main article: Two-dimensional infrared spectroscopy analysis Two-dimensional infrared correlation spectroscopy analysis is the application of 2D correlation analysis on infrared spectra. By extending the spectral information of a perturbed sample, spectral analysis is simplified and resolution is enhanced. The 2D synchronous and 2D asynchronous spectra represent a graphical overview of the spectral changes due to a perturbation (such as a changing concentration or changing temperature) as well as the relationship between the spectral changes at two different wavenumbers. Main article: Nonlinear two-dimensional infrared spectroscopy Nonlinear two-dimensional infrared spectroscopy[2][3] is the infrared version of correlation spectroscopy. Nonlinear two-dimensional infrared spectroscopy is a technique that has come available with the development of femtosecond infrared laser pulses. In this experiment first a set of pump pulses are applied to the sample. This is followed by a waiting time, where the system is allowed to relax. The waiting time typically lasts from zero to several picoseconds and the duration can be controlled with a resolution of tens of femtoseconds. A probe pulse is then applied resulting in the emission of a signal from the sample. The nonlinear two-dimensional infrared spectrum is a two-dimensional correlation plot of the frequency 1 that was excited by the initial pump pulses and the frequency 3 excited by the probe pulse after the waiting time. This allows the observation of coupling between different vibrational modes. Because of its extremely high time resolution it can be used to follow changes in molecular configurations taking place on a picosecond timescale. It is still a largely unexplored technique and is becoming increasingly popular for fundamental research. Like in two-dimensional nuclear magnetic resonance (2DNMR) spectroscopy this technique spreads the spectrum in two dimensions and allow for the observation of cross peaks that contain information on the coupling between different modes. In contrast to 2DNMR nonlinear two-dimensional infrared spectroscopy also involve the excitation to overtones. These excitations result in excited state absorption peaks located below the diagonal and cross peaks. In 2DNMR two distinct techniques, COSY and NOESY, are

frequently used. The cross peaks in the first are related to the scalar coupling, while in the later they are related to the spin transfer between different nuclei. In nonlinear twodimensional infrared spectroscopy analogs have been drawn to these 2DNMR techniques. Nonlinear two-dimensional infrared spectroscopy with zero waiting time corresponds to COSY and nonlinear two-dimensional infrared spectroscopy with finite waiting time allowing vibrational population transfer corresponds to NOESY. The COSY variant of nonlinear two-dimensional infrared spectroscopy has been used for determination of the secondary structure content proteins.[4]

[edit] See also


Infrared spectroscopy correlation table Fourier transform spectroscopy Near infrared spectroscopy Vibrational spectroscopy Rotational spectroscopy Spectroscopy Quantum vibration Raman spectroscopy Frequency comb

[edit] References
1. ^ Lau, W.S. (1999). Infrared characterization for microelectronics. World Scientific. 2. ^ P. Hamm, M. H. Lim, R. M. Hochstrasser (1998). "Structure of the amide I band of peptides measured by femtosecond nonlinear-infrared spectroscopy". J. Phys. Chem. B 102: 6123. 3. ^ S. Mukamel (2000). "Multidimensional Fentosecond Correlation Spectroscopies of Electronic and Vibrational Excitations". Annu. Rev. Phys. Chem. 51: 691. 4. ^ N. Demirdven, C. M. Cheatum, H. S. Chung, M. Khalil, J. Knoester, A. Tokmakoff (2004). "Two-dimensional infrared spectroscopy of antiparallel betasheet secondary structure". J. Am. Chem. Soc. 126: 7981.

[edit] External links

The Science of Spectroscopy - supported by NASA. Spectroscopy education wiki and films - introduction to light, its uses in NASA, space science, astronomy, medicine & health, environmental research, and consumer products. A useful gif animation of different vibrational modes: here Infrared spectroscopy for organic chemists Spectral Calculator - instructive simulation tool. High resolution line-by-line molecular spectrum modelling using NASA algorithms.

vde

Spectroscopy[hide]
Atomic spectroscopy Emission spectroscopy Electron spin resonance Fluorescence spectroscopy Gamma spectroscopy Infrared spectroscopy Laser induced breakdown spectroscopy Mbauer spectroscopy Nuclear magnetic resonance spectroscopy Raman spectroscopy Resonance enhanced multiphoton ionization Rotational spectroscopy Terahertz spectroscopy Ultravioletvisible spectroscopy Vibrational spectroscopy X-ray spectroscopy

Retrieved from "http://en.wikipedia.org/wiki/Infrared_spectroscopy" Category: Spectroscopy

You might also like