Sie sind auf Seite 1von 36

UV-VISIBLE ABSORPTION SPECTRA

This page explains what happens when organic compounds absorb UV or visible light, and why the wavelength of light absorbed varies from compound to compound.
Important: If you have come direct to this page from a search engine, you should be aware that it follows on from an introductory page explaining some essential bonding for UV-visible spectrometry. You need to read this before you go on with this page.

What happens when light is absorbed by molecules?


Promotion of electrons When we were talking about the various sorts of orbitals present in organic compounds on the introductory page (see above), you will have come across this diagram showing their relative energies:

Remember that the diagram isn't intended to be to scale - it just shows the relative placing of the different orbitals. When light passes through the compound, energy from the light is used to promote an electron from a bonding or non-bonding orbital into one of the empty anti-bonding orbitals.

Important: If you don't know exactly what I mean by bonding, non-bonding and antibonding orbitals, or don't remember the diagram, go back and read the introductory page again.

The possible electron jumps that light might cause are:

In each possible case, an electron is excited from a full orbital into an empty anti-bonding orbital. Each jump takes energy from the light, and a big jump obviously needs more energy than a small one. Each wavelength of light has a particular energy associated with it. If that particular amount of energy is just right for making one of these energy jumps, then that wavelength will be absorbed - its energy will have been used in promoting an electron. We need to work out what the relationship is between the energy gap and the wavelength absorbed. Does, for example, a bigger energy gap mean that light of a lower wavelength will be absorbed - or what? It is easier to start with the relationship between the frequency of light absorbed and its energy:

You can see that if you want a high energy jump, you will have to absorb light of a higher frequency. The greater the frequency, the greater the energy. That's easy - but unfortunately UV-visible absorption spectra are always given using wavelengths of light rather than frequency. That means that you need to know the relationship between wavelength and frequency.

You can see from this that the higher the frequency is, the lower the wavelength is. So . . . If you have a bigger energy jump, you will absorb light with a higher frequency - which is the same as saying that you will absorb light with a lower wavelength.

Important summary The larger the energy jump, the lower the wavelength of the light absorbed.
Note: It is obviously better if you can work this out in case you forget it, but you may feel that it is a lot easier just to learn the last statement. However you do it, you must

be confident about this relationship. If you need more help with sorting out these relationships, you will find them discussed more slowly on a page in this section about electromagnetic radiation. If you choose to follow this link, use the BACK button on your browser to return to this page.

Some jumps are more important than others for absorption spectrometry An absorption spectrometer works in a range from about 200 nm (in the near ultra-violet) to about 800 nm (in the very near infra-red). Only a limited number of the possible electron jumps absorb light in that region.
Note: If you are interested, there is a description of how a double beam absorption spectrometer works on another page in this section. It isn't necessary to know about this in order to understand the rest of this page.

Look again at the possible jumps. This time, the important jumps are shown in black, and a less important one in grey. The grey dotted arrows show jumps which absorb light outside the region of the spectrum we are working in.

Remember that bigger jumps need more energy and so absorb light with a shorter wavelength. The jumps shown with grey dotted arrows absorb UV light

of wavelength less that 200 nm. The important jumps are:


y y y

from pi bonding orbitals to pi anti-bonding orbitals; from non-bonding orbitals to pi anti-bonding orbitals; from non-bonding orbitals to sigma anti-bonding orbitals.

That means that in order to absorb light in the region from 200 - 800 nm (which is where the spectra are measured), the molecule must contain either pi bonds or atoms with non-bonding orbitals. Remember that a non-bonding orbital is a lone pair on, say, oxygen, nitrogen or a halogen. Groups in a molecule which absorb light are known as chromophores.
Note: A Canadian university site describes "chromophore" as "one of those useful but sloppy words whose meaning depends somewhat on the context." As you will find shortly, highly delocalised systems in a molecule are often responsible for absorbing light in the visible region and making the compound look coloured. The delocalisation may involve several different types of group - benzene rings, carboncarbon double bonds, carbon-oxygen double bonds, lone pairs on nitrogen or oxygen and so on. Some people talk as if the whole of the delocalised system was the chromophore; others tend to think on terms of the contributions of individual parts of the system. It seems to me to be more logical to think in terms of the whole system because it is electron jumps within the system as a whole that cause absorption. If your examiners take a different view, however, obviously you should go with what they want. The only way you will find that out is to look at recent exam papers and mark schemes. If you are a UK A level (or equivalent) student follow this link to the syllabuses page to find out how to get hold of these if you haven't already got them.

What does an absorption spectrum look like The diagram below shows a simple UV-visible absorption spectrum for buta1,3-diene - a molecule we will talk more about later. Absorbance (on the vertical axis) is just a measure of the amount of light absorbed. The higher the value, the more of a particular wavelength is being absorbed.

You will see that absorption peaks at a value of 217 nm. This is in the ultraviolet and so there would be no visible sign of any light being absorbed - buta1,3-diene is colourless. You read the symbol on the graph as "lambda-max". In buta-1,3-diene, CH2=CH-CH=CH2, there are no non-bonding electrons. That means that the only electron jumps taking place (within the range that the spectrometer can measure) are from pi bonding to pi anti-bonding orbitals.
Note: The spectrum shown is only a simplified sketch graph - it doesn't pretend to any great accuracy! If you are really wide-awake you might wonder why the graph looks like it does with a broad absorption peak rather than a single line at 217 nm. A jump from a pi bonding orbital to a pi anti-bonding orbital ought to have a fixed energy and therefore absorb a fixed wavelength. The compound is in fact absorbing over a whole range of wavelengths suggesting a whole range of energy jumps. This problem arises because rotations and vibrations in the molecule are continually changing the energies of the orbitals - and that, of course, means that the gaps between them are continually changing as well. The result is that absorption takes place over a range of wavelengths rather than at one fixed one. You don't need to worry about this at this level.

A chromophore producing two peaks A chromophore such as the carbon-oxygen double bond in ethanal, for example, obviously has pi electrons as a part of the double bond, but also has lone pairs on the oxygen atom.

That means that both of the important absorptions from the last energy diagram are possible.
Note: If you aren't too sure about the bonding in a carbon-oxygen double bond, it would be worth following this link before you go on. Use the BACK button on your browser to return to this page when you are ready.

You can get an electron excited from a pi bonding to a pi anti-bonding orbital, or you can get one excited from an oxygen lone pair (a non-bonding orbital) into a pi anti-bonding orbital.

Note: Before you read on, work out which of these will absorb light with the longer wavelength. Try it! If you can work this out for yourself, you have cracked one of the most difficult things about this topic.

The non-bonding orbital has a higher energy than a pi bonding orbital. That means that the jump from an oxygen lone pair into a pi anti-bonding orbital needs less energy. That means it absorbs light of a lower frequency and therefore a higher wavelength. Ethanal can therefore absorb light of two different wavelengths:
y y

the pi bonding to pi anti-bonding absorption peaks at 180 nm; the non-bonding to pi anti-bonding absorption peaks at 290 nm.

Both of these absorptions are in the ultra-violet, but most spectrometers won't pick up the one at 180 nm because they work in the range from 200 - 800 nm.

The importance of conjugation and delocalisation in what wavelength is absorbed Consider these three molecules:

Ethene contains a simple isolated carbon-carbon double bond, but the other two have conjugated double bonds. In these cases, there is delocalisation of the pi bonding orbitals over the whole molecule.
Important: If you don't know understand about conjugation, go back and read that section of the introductory page again.

Now look at the wavelengths of the light which each of these molecules absorbs. molecule ethene buta-1,3-diene hexa-1,3,5-triene wavelength of maximum absorption (nm) 171 217 258

All of the molecules give similar UV-visible absorption spectra - the only difference being that the absorptions move to longer and longer wavelengths as the amount of delocalisation in the molecule increases. Why is this? You can actually work out what must be happening.
y y

The maximum absorption is moving to longer wavelengths as the amount of delocalisation increases. Therefore maximum absorption is moving to shorter frequencies as

y y

the amount of delocalisation increases. Therefore absorption needs less energy as the amount of delocalisation increases. Therefore there must be less energy gap between the bonding and anti-bonding orbitals as the amount of delocalisation increases.

. . . and that's what is happening. Compare ethene with buta-1,3-diene. In ethene, there is one pi bonding orbital and one pi anti-bonding orbital. In buta-1,3-diene, there are two pi bonding orbitals and two pi anti-bonding orbitals. This is all discussed in detail on the introductory page that you should have read.

The highest occupied molecular orbital is often referred to as the HOMO - in these cases, it is a pi bonding orbital. The lowest unoccupied molecular orbital (the LUMO) is a pi anti-bonding orbital. Notice that the gap between these has fallen. It takes less energy to excite an electron in the buta-1,3-diene case than with ethene. In the hexa-1,3,5-triene case, it is less still.

Note: In this case, you will have to work out for yourself why there are 3 bonding and 3 anti-bonding pi orbitals in the hexa-1,3,5-triene. If you are interested, you can work it out in the same way that I have used for buta-1,3-diene on the introductory page. It isn't essential that you do this - all that matters is that you can see the pattern - the more the delocalisation, the longer the wavelength absorbed. If you choose to follow this link, use the BACK button on your browser to come back here again later.

If you extend this to compounds with really massive delocalisation, the wavelength absorbed will eventually be high enough to be in the visible region of the spectrum, and the compound will then be seen as coloured. A good example of this is the orange plant pigment, beta-carotene - present in carrots, for example.

Why is beta-carotene orange? Beta-carotene has the sort of delocalisation that we've just been looking at, but on a much greater scale with 11 carbon-carbon double bonds conjugated together. The diagram shows the structure of beta-carotene with the alternating double and single bonds shown in red.

The more delocalisation there is, the smaller the gap between the highest energy pi bonding orbital and the lowest energy pi anti-bonding orbital. To promote an electron therefore takes less energy in beta-carotene than in the cases we've looked at so far - because the gap between the levels is less. Remember that less energy means a lower frequency of light gets absorbed and that's equivalent to a longer wavelength. Beta-carotene absorbs throughout the ultra-violet region into the violet - but particularly strongly in the visible region between about 400 and 500 nm with a peak about 470 nm. If you have read the page in this section about electromagnetic radiation, you might remember that the wavelengths associated with the various colours are approximately: colour region wavelength (nm) violet blue cyan green yellow orange red 380 - 435 435 - 500 500 - 520 520 - 565 565 - 590 590 - 625 625 - 740

So if the absorption is strongest in the violet to cyan region, what colour will you actually see? It is tempting to think that you can work it out from the colours that are left - and in this particular case, you wouldn't be far wrong. Unfortunately, it isn't as simple as that! Sometimes what you actually see is quite unexpected. Mixing different wavelengths of light doesn't give you the same result as mixing paints or other pigments. You can, however, sometimes get some estimate of the colour you would see using the idea of complementary colours.

Complementary colours If you arrange some colours in a circle, you get a "colour wheel". The diagram shows one possible version of this. An internet search will throw up many different versions!

Colours directly opposite each other on the colour wheel are said to be complementary colours. Blue and yellow are complementary colours; red and cyan are complementary; and so are green and magenta. Mixing together two complementary colours of light will give you white light.
Beware: That is NOT the same as mixing together paint colours. If you mix yellow and blue paint you don't get white paint. Is this confusing? YES!

What this all means is that if a particular colour is absorbed from white light, what your eye detects by mixing up all the other wavelengths of light is its complementary colour. In the beta-carotene case, the situation is more confused because you are absorbing such a range of wavelengths. However, if you think of the peak absorption running from the blue into the cyan, it would be reasonable to think of the colour you would see as being opposite that where yellow runs into red in other words, orange. You will find the colours a bit more clear-cut in the other two examples we'll talk about below.
Note: If you are interested in understanding the relationship between colour absorbed and colour seen (beyond the very basic description above), find your way to lesson 2 ("Color and Vision") of "Light Waves and Vision" on The Physics Classroom. I'm not giving a direct link to those pages, because that site is still developing and it is safer to give a link to the front page of the site. This is the most understandable explanation I have found anywhere on the web.

Applying this to the colour changes of two indicators


Phenolphthalein You have probably used phenolphthalein as an acid-base indicator, and will know that it is colourless in acidic conditions and magenta (bright pink) in an alkaline solution. How is this colour change related to changes in the molecule? The structures of the two differently coloured forms are:

Both of these absorb light in the ultra-violet, but the one on the right also absorbs in the visible with a peak at 553 nm. The molecule in acid solution is colourless because our eyes can't detect the fact that some light is being absorbed in the ultra-violet. However, our eyes do detect the absorption at 553 nm produced by the form in alkaline solution. 553 nm is in the green region of the spectrum. If you look back at the colour wheel, you will find that the complementary colour of green is magenta - and that's the colour you see. So why does the colour change as the structure changes? What we have is a shift to absorption at a higher wavelength in alkaline solution. As we've already seen, a shift to higher wavelength is associated with a greater degree of delocalisation. Here is a modified diagram of the structure of the form in acidic solution - the colourless form. The extent of the delocalisation is shown in red.

Notice that there is delocalisation over each of the three rings - extending out over the carbon-oxygen double bond, and to the various oxygen atoms because of their lone pairs. But the delocalisation doesn't extend over the whole molecule. The carbon atom in the centre with its four single bonds prevents the three delocalised regions interacting with each other. Now compare that with the magenta form:

The rearrangement now lets the delocalisation extend over the entire ion. This greater delocalisation lowers the energy gap between the highest occupied molecular orbital and the lowest unoccupied pi anti-bonding orbital. It needs less energy to make the jump and so a longer wavelength of light is absorbed. Remember: Increasing the amount of delocalisation shifts the absorption peak to a higher wavelength.
Note: Don't panic at the thought that you might have to draw these structures in an exam. At UK A level (and its equivalent) standard, it won't happen. However, if you were given the structures of the two forms of phenolphthalein and asked to explain the colour change, that could make a good question (depending, of course, on what exactly your syllabus expects you to be able to do).

Methyl orange You will know that methyl orange is yellow in alkaline solutions and red in

acidic ones. The structure in alkaline solution is:

In acid solution, a hydrogen ion is (perhaps unexpectedly) picked up on one of the nitrogens in the nitrogen-nitrogen double bond.

This now gets a lot more complicated! The positive charge on the nitrogen is delocalised (spread around over the structure) - especially out towards the right-hand end of the molecule as we've written it. The normally drawn structure for the red form of methyl orange is . . .

But this can be seriously misleading as regards the amount of delocalisation in the structure for reasons discussed below (after the red warning box) if you are interested.
Important: If you have read the introductory page, you will know that I am uncertain about whether the delocalisation extends out over the sulphonate group or not. That doesn't affect the rest of this argument in any way. This group isn't changed by the addition of a hydrogen ion. All we are actually interested in is the effect on the delocalisation over the rest of the molecule - the two benzene rings and the two nitrogen-containing groups.

So which is the more delocalised structure - red or yellow? Let's work backwards from the absorption spectra to see if that helps. The yellow form has an absorption peak at about 440 nm. That's in the blue

region of the spectrum, and the complementary colour of blue is yellow. That's exactly what you would expect. The red form has an absorption peak at about 520 nm. That's at the edge of the cyan region of the spectrum, and the complementary colour of cyan is red. Again, there's nothing unexpected here. Notice that the change from the yellow form to the red form has produced an increase in the wavelength absorbed. An increase in wavelength suggests an increase in delocalisation. That means that there must be more delocalisation in the red form than in the yellow one. Why? For the purposes of UK A level (and its equivalents), that's probably unanswerable. However, if you are interested, there's a possible answer below ...
Warning! The rest of this page is going to get seriously difficult. If you are a UK A level (or equivalent) student, it goes well beyond the level of understanding that you are likely to need. The reason for including it is to try to remove the impression that the red form is less delocalised than the yellow one because of the way the structure is usually drawn. There was even a question on an OCR A level paper concerning methyl orange that told you (wrongly!) that there was less delocalisation in the red form than in the yellow one. If you are a UK OCR student, please don't use Q3 (b)(ii) on paper 2815-04 June 2005 for practice - it is unanswerable because the whole basis of the question is wrong.

Here again is the structure of the yellow form:

Delocalisation will extend over most of the structure - out as far as the lone pair on the right-hand nitrogen atom. If you use the normally written structure for the red form, the delocalisation seems to be broken in the middle - the pattern of alternating single and double

bonds seems to be lost.

But that is to misunderstand what this last structure represents.

Canonical forms If you draw the two possible Kekul structures for benzene, you will know that the real structure of benzene isn't like either of them. The real structure is somewhere between the two - all the bonds are identical and somewhere between single and double in character. That's because of the delocalisation in benzene.

The two structures are known as canonical forms, and they can each be thought of as adding some knowledge to the real structure. For example, the bond drawn at the top right of the molecule is neither truly single or double, but somewhere in between. Similarly with all the other bonds.
Note: If you haven't come across canonical forms as a way of representing delocalisation, it is important that you don't imagine that the molecule is rapidly flipping from one structure to another. The double-headed arrows mean something different. A mule is a hybrid of a donkey and a horse. In this notation, you could represent a mule by writing donkey and horse connected by a double-headed arrow. Neitherdonkey nor horse accurately represents what a mule looks like, but with a bit of imagination you could build up a fairly good picture of a mule by combining together the characteristics of both donkey and horse. But a mule obviously doesn't spend its time rapidly changing back and forth between being a donkey and a horse!

The two structures we've previously drawn for the red form of methyl orange are also canonical forms - two out of lots of forms that could be drawn for this structure. We could represent the delocalised structure by:

These two forms can be thought of as the result of electron movements in the structure, and curly arrows are often used to show how one structure can lead to the other.

In reality, the electrons haven't shifted fully either one way or the other. Just as in the benzene case, the actual structure lies somewhere in between these. You must also realise that drawing canonical forms has no effect on the underlying geometry of the structure. Bond types or lengths or angles don't change in the real structure. For example, the lone pairs on the nitrogen atoms shown in the last diagram are both involved with the delocalisation. For this to happen all the bonds around these nitrogens must be in the same plane, with the lone pair sticking up so that it can overlap sideways with orbitals on the next-door atoms. The fact that in each of the two canonical forms one of these nitrogens is shown as if it had an ammonia-like arrangement of the bonds is potentially misleading and makes it look as if the delocalisation is broken. The problem is that there is no easy way of representing a complex delocalised structure in simple structural diagrams. It is bad enough with benzene - with something as complicated as methyl orange any method just

leads to possible confusion if you aren't used to working with canonical forms. It gets even more complicated! If you were doing this properly there would be a host of other canonical forms with different arrangements of double and single bonds and with the positive charge located at various places around the rings and on the other nitrogen atom. The real structure can't be represented properly by any one of this multitude of canonical forms, but each gives a hint of how the delocalisation works. If we take the two forms we have written as perhaps the two most important ones, it suggests that there is delocalisation of the electrons over the whole structure, but that electron density is a bit low around the two nitrogens carrying the positive charge on one canonical form or the other.

So why is the red form more delocalised than the yellow one? Finally, we get around to an attempt at an explanation as to why the delocalisation is greater in the red form of methyl orange in acid solution than in the yellow one in alkaline solution. The answer may lie in the fact that the lone pair on the nitrogen at the righthand end of the structure as we've drawn it is more fully involved in the delocalisation in the red form. The canonical form with the positive charge on that nitrogen suggests a significant movement of that lone pair towards the rest of the molecule. Doesn't the same thing happen to the lone pair on the same nitrogen in the yellow form of methyl orange? Not to the same extent. Any canonical form that you draw in which that happens produces another negatively charged atom somewhere in the rest of the structure. Separating negative and positive charges like this is energetically unfavourable. In the red form, we aren't producing a new separation of charge - just shifting a positive charge around the structure.

Please remember: This is high-level stuff. To really understand it you need to have met canonical forms before and practised drawing them. This is degree level chemistry and not A level (or its equivalent). You do NOT need any of this (from below the red warning) unless perhaps you are reading this as a university student.

USING UV-VISIBLE ABSORPTION SPECTRA

This page takes a brief look at how UV-visible absorption spectra can be used to help identify compounds and to measure the concentrations of coloured solutions. It assumes that you know how these spectra arise, and know what is meant by terms such as absorbance, molar absorptivity and lambda-max. You also need to be familiar with the Beer-Lambert Law.
Important: If you don't know about these things, this page is a waste of your time! Explore the rest of the UV-visible spectroscopy menu before you go on.

Using UV-absorption spectra to help identify organic compounds If you have worked through the rest of this section, you will know that the wavelength of maximum absorption (lambda-max) depends on the presence of particular chromophores (light-absorbing groups) in a molecule. For example, on another page you will have come across the fact that a simple carbon-carbon double bond (for example in ethene) has a maximum absorption at 171 nm. The two conjugated double bonds in buta-1,3-diene have a maximum absorption at a longer wavelength of 217 nm. We also talked about the two peaks in the spectrum of ethanal (containing a simple carbon-oxygen double bond) at 180 and 290 nm. In carefully chosen simple cases (which is all you will get at this level), if you compared the peaks on a given UV-visible absorption spectrum with a list of known peaks, it would be fairly easy to pick out some structural features of an unknown molecule. Lists of known peaks often include molar absorptivity values as well. That might help you to be even more sure. For example (again using the simple carbon-oxygen double bond), data shows that the peak at 290 has a molar absorptivity of only 15, compared with the one at 180 of 10000. If your spectrum showed a very large peak at 180, and an extremely small one at 290, that just adds to your certainty. Any question set at this level is going to be so trivial, and so obvious, that it isn't worth spending any more time on this. Let's look at something a bit more

complicated!

Using UV-absorption spectra to find concentrations You should remember the Beer-Lambert Law:

The expression on the left of the equation is known as the absorbance of the solution and is measured by a spectrometer. The equation is sometimes written in terms of that absorbance.

The symbol epsilon is the molar absorptivity of the solution.


Note: You will find this all explained in more detail on the page about the Beer-Lambert Law.

Finding concentration using the molar absorptivity If you know the molar absorptivity of a solution at a particular wavelength, and you measure the absorbance of the solution at that wavelength, it is easy to calculate the concentration. The only other variable in the expression above is the length of the solution. That's easy to measure and, in fact, the cell containing the solution may well have been manufactured with a known length

of 1 cm. For example, let's suppose you have a solution in a cell of length 1 cm. You measure the absorbance of the solution at a particular wavelength using a spectrometer. The value is 1.92. You find a value for molar absorptivity in a table of 19400 for that wavelength. Substituting those values:

Notice what a very low concentration can be measured provided you are working with a substance with a very high molar absorptivity. This method, of course, depends on you having access to an accurate value of molar absorptivity. It also assumes that the Beer-Lambert Law works over the whole concentration range (not true!). It is much better to measure the concentration by plotting a calibration curve. It saves doing any calculations for one thing!

Finding concentration by plotting a calibration curve Doing it this way you don't have to rely on a value of molar absorptivity, the reliability of the Beer-Lambert Law, or even know the dimensions of the cell containing the solution. What you do is make up a number of solutions of the compound you are investigating - each of accurately known concentration. Those concentrations should bracket the concentration you are trying to find - some less concentrated; some more concentrated. With coloured solutions, this isn't a problem. You would just make up some solutions which are a bit lighter and some a bit darker in colour. For each solution, you measure the absorbance at the wavelength of strongest

absorption - using the same container for each one. Then you plot a graph of that absorbance against concentration. This is a calibration curve. According to the Beer-Lambert Law, absorbance is proportional to concentration, and so you would expect a straight line. That is true as long as the solutions are dilute, but the Law breaks down for solutions of higher concentration, and so you might get a curve under these circumstances. As long as you are working from values either side of the one you are trying to find, that isn't a problem. Having drawn a best fit line, the calibration curve will probably look something like the next diagram. (I've drawn it as a straight line because it is easier for me to draw than a curve(!), and it's what you will probably get if you are working with really dilute solutions. But if it turns out to be a curve, so be it!)

Notice that no attempt has been made to force the line back through the origin. If the Beer-Lambert Law worked perfectly, it would pass through the origin, but you can't guarantee that it is working properly at the concentrations you are using. Now all you have to do is to measure the absorbance of the solution with the unknown concentration at the same wavelength. If, for example, it had an absorbance of 0.600, you can just read the corresponding concentration from the graph as below.

Visible and Ultraviolet Spectroscopy


1. Background
An obvious difference between certain compounds is their color. Thus, quinone is yellow; chlorophyll is green; the 2,4dinitrophenylhydrazone derivatives of aldehydes and ketones range in color from bright yellow to deep red, depending on double bond conjugation; and aspirin is colorless. In this respect the human eye is functioning as a spectrometer analyzing the light reflected from the surface of a solid or passing through a liquid. Although we see sunlight (or white light) as uniform or homogeneous in color, it is actually composed of a broad range of radiation wavelengths in the ultraviolet (UV), visible and infrared (IR) portions of the spectrum. As shown on the right, the component colors of the visible portion can be separated by passing sunlight through a prism, which acts to bend the light in differing degrees according to wavelength. Electromagnetic radiation such as visible light is commonly treated as a wave phenomenon, characterized by a wavelength or frequency. Wavelength is defined on the left below, as the distance between adjacent peaks (or troughs), and may be designated in meters, centimeters or nanometers (10-9 meters). Frequency is the number of wave cycles that travel past a fixed point per unit of time, and is usually given in cycles per second, or hertz (Hz). Visible wavelengths cover a range from approximately 400 to 800 nm. The longest visible wavelength is red and the shortest is violet. Other common colors of the spectrum, in order of decreasing wavelength, may be remembered by the mnemonic: ROY G BIV. The wavelengths of what we perceive as particular colors in the visible portion of the spectrum are displayed and listed below. In horizontal diagrams, such as the one on the bottom left, wavelength will increase on moving from left to right.
y y y y y

Violet: 400 - 420 nm Indigo: 420 - 440 nm Blue: 440 - 490 nm Green: 490 - 570 nm Yellow: 570 - 585 nm

y y

Orange: 585 - 620 nm Red: 620 - 780 nm

When white light passes through or is reflected by a colored substance, a characteristic portion of the mixed wavelengths is absorbed. The remaining light will then assume the complementary color to the wavelength(s) absorbed. This relationship is demonstrated by the color wheel shown on the right. Here, complementary colors are diametrically opposite each other. Thus, absorption of 420-430 nm light renders a substance yellow, and absorption of 500-520 nm light makes it red. Green is unique in that it can be created by absoption close to 400 nm as well as absorption near 800 nm. Early humans valued colored pigments, and used them for decorative purposes. Many of these were inorganic minerals, but several important organic dyes were also known. These included the crimson pigment, kermesic acid, the blue dye, indigo, and the yellow saffron pigment, crocetin. A rare dibromo-indigo derivative, punicin, was used to color the robes of the royal and wealthy. The deep orange hydrocarbon carotene is widely distributed in plants, but is not sufficiently stable to be used as permanent pigment, other than for food coloring. A common feature of all these colored compounds, displayed below, is a system of extensively conjugated pi-electrons.

2. The Electromagnetic Spectrum


The visible spectrum constitutes but a small part of the total radiation spectrum. Most of the radiation that surrounds us cannot be seen, but can be detected by dedicated sensing instruments. This electromagnetic spectrum ranges from very short wavelengths (including gamma and x-rays) to very long wavelengths (including microwaves and broadcast radio waves). The following chart displays many of the important regions of this spectrum, and demonstrates the inverse relationship between wavelength and frequency (shown in the top equation below the chart).

The energy associated with a given segment of the spectrum is proportional to its frequency. The bottom equation describes this relationship, which provides the energy carried by a photon of a given wavelength of radiation.

To obtain specific frequency, wavelength and energy values use this calculator.

3. UV-Visible Absorption Spectra


To understand why some compounds are colored and others are not, and to determine the relationship of conjugation to color, we must make accurate measurements of light absorption at different wavelengths in and near the visible part of the spectrum. Commercial optical spectrometers enable such experiments to be conducted with ease, and usually survey both the near ultraviolet and visible portions of the spectrum. For a description of a UV-Visible spectrometer Click Here. The visible region of the spectrum comprises photon energies of 36 to 72 kcal/mole, and the near ultraviolet region, out to 200 nm, extends this energy range to 143 kcal/mole. Ultraviolet radiation having wavelengths less than 200 nm is difficult to handle, and is seldom used as a routine tool for structural analysis.

The energies noted above are sufficient to promote or excite a molecular electron to a higher energy orbital. Consequently, absorption spectroscopy carried out in this region is sometimes called "electronic spectroscopy". A diagram showing the various kinds of electronic excitation that may occur in organic molecules is shown on the left. Of the six transitions outlined, only the two lowest energy ones (left-most, colored blue) are achieved by the energies available in the 200 to 800 nm spectrum. As a rule, energetically favored electron promotion will be from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO), and the resulting species is called an excited state. For a review of molecular orbitals click here. When sample molecules are exposed to light having an energy that matches a possible electronic transition within the molecule, some of the light energy will be absorbed as the electron is promoted to a higher energy orbital. An optical spectrometer records the wavelengths at which absorption occurs, together with the degree of absorption at each wavelength. The resulting spectrum is presented as a graph of absorbance (A) versus wavelength, as in the isoprene spectrum shown below. Since isoprene is colorless, it does not absorb in the visible part of the spectrum and this region is not displayed on the graph. Absorbance usually ranges from 0 (no absorption) to 2 (99% absorption), and is precisely defined in context with spectrometer operation. Because the absorbance of a sample will be proportional to the number of absorbing molecules in the spectrometer light beam (e.g. their molar concentration in the sample tube), it is necessary to correct the absorbance value for this and other operational factors if the spectra of different compounds are to be compared in a meaningful way. The corrected absorption value is called "molar absorptivity", and is particularly useful when comparing the spectra of different compounds and determining the relative strength of light absorbing functions (chromophores). Molar absorptivity ( ) is defined as: Molar Absorptivity, A/cl (where A= absorbance, c = sample concentration in moles/liter & l = length of light path through the sample in cm.)

If the isoprene spectrum on the right was obtained from a dilute hexane solution (c = 4 * 10-5 moles per liter) in a 1 cm sample cuvette, a simple calculation using the above formula indicates a molar absorptivity of 20,000 at the maximum absorption wavelength. Indeed the entire vertical absorbance scale may be changed to a molar absorptivity scale once this information about the sample is in hand. Clicking on the spectrum will display this change in units.

Chromophore C=C C C C=O N=O C-X X=Br X=I

Example Ethene 1-Hexyne Ethanal Nitromethane Methyl bromide Methyl Iodide

Excitation
__

max

, nm

> > > > > > > >

* * * * * * * *

171 180 290 180 275 200 205 255

__

n n n n

__

__

__

__

__ __

From the chart above it should be clear that the only molecular moieties likely to absorb light in the 200 to 800 nm region are pi-electron functions and hetero atoms having non-bonding valence-shell electron pairs. Such light absorbing groups are referred to as chromophores. A list of some simple chromophores and their light absorption characteristics is provided on the left above. The oxygen non-bonding electrons in alcohols and ethers do not give rise to absorption above 160 nm. Consequently, pure alcohol and ether solvents may be used for spectroscopic studies. The presence of chromophores in a molecule is best documented by UV-Visible spectroscopy, but the failure of most instruments to provide absorption data for wavelengths below 200 nm makes the detection of isolated chromophores problematic. Fortunately, conjugation generally moves the absorption maxima to longer wavelengths, as in the case of isoprene, so conjugation becomes the major structural feature identified by this technique. Molar absorptivities may be very large for strongly absorbing chromophores (>10,000) and very small if absorption is weak (10 to 100). The magnitude of reflects both the size of the chromophore and the probability that light of a given wavelength will be absorbed when it strikes the chromophore. For further discussion of this topic Click Here.

4. The Importance of Conjugation


A comparison of the absorption spectrum of 1-pentene, max = 178 nm, with that of isoprene (above) clearly demonstrates the importance of chromophore conjugation. Further evidence of this effect is shown below. The spectrum on the left illustrates that conjugation of double and triple bonds also shifts the absorption maximum to longer wavelengths. From the polyene spectra displayed in the center diagram, it is clear that each additional double bond in the conjugated pi-electron system shifts the absorption maximum about 30 nm in the same direction. Also, the molar absorptivity ( ) roughly doubles with each new conjugated double bond. Spectroscopists use the terms defined in the table on the right when describing shifts in absorption. Thus, extending conjugation generally results in bathochromic and hyperchromic shifts in absorption. The appearance of several absorption peaks or shoulders for a given chromophore is common

for highly conjugated systems, and is often solvent dependent. This fine structure reflects not only the different conformations such systems may assume, but also electronic transitions between the different vibrational energy levels possible for each electronic state. Vibrational fine structure of this kind is most pronounced in vapor phase spectra, and is increasingly broadened and obscured in solution as the solvent is changed from hexane to methanol.

Terminology for Absorption Shifts


Nature of Shift To Longer Wavelength To Shorter Wavelength To Greater Absorbance To Lower Absorbance Descriptiv e Term Bathochro mic Hypsochro mic Hyperchro mic Hypochrom ic

To understand why conjugation should cause bathochromic shifts in the absorption maxima of chromophores, we need to look at the relative energy levels of the pi-orbitals. When two double bonds are conjugated, the four p-atomic orbitals combine to generate four pi-molecular orbitals (two are bonding and two are antibonding). This was described earlier in the section concerning diene chemistry. In a similar manner, the three double bonds of a conjugated triene create six pi-molecular orbitals, half bonding and half antibonding. The energetically most favorable __ > * excitation occurs from the highest energy bonding pi-orbital (HOMO) to the lowest energy antibonding pi-orbital (LUMO). The following diagram illustrates this excitation for an isolated double bond (only two pi-orbitals) and, on clicking the diagram, for a conjugated diene and triene. In each case the HOMO is colored blue and the LUMO is colored magenta. Increased conjugation brings the HOMO and LUMO orbitals closer together. The energy ( E) required to effect the electron promotion is therefore less, and the wavelength that provides this energy is increased correspondingly (remember = h c/ E ).

Examples of __> * Excitation Click on the Diagram to Advance

Many other kinds of conjugated pi-electron systems act as chromophores and absorb light in the 200 to 800 nm region. These include unsaturated aldehydes and ketones and aromatic ring compounds. A few examples are displayed below. The spectrum of the unsaturated ketone (on the left) illustrates the advantage of a logarithmic display of molar absorptivity. The __> * absorption located at 242 nm is very strong, with an = 18,000. The weak n __> * absorption near 300 nm has an = 100.

Benzene exhibits very strong light absorption near 180 nm ( > 65,000) , weaker absorption at 200 nm ( = 8,000) and a group of much weaker bands at 254 nm ( = 240). Only the last group of absorptions are completely displayed because of the 200 nm cut-off characteristic of most

spectrophotometers. The added conjugation in naphthalene, anthracene and tetracene causes bathochromic shifts of these absorption bands, as displayed in the chart on the left below. All the absorptions do not shift by the same amount, so for anthracene (green shaded box) and tetracene (blue shaded box) the weak absorption is obscured by stronger bands that have experienced a greater red shift. As might be expected from their spectra, naphthalene and anthracene are colorless, but tetracene is orange.

The spectrum of the bicyclic diene (above right) shows some vibrational fine structure, but in general is similar in appearance to that of isoprene, shown above. Closer inspection discloses that the absorption maximum of the more highly substituted diene has moved to a longer wavelength by about 15 nm. This "substituent effect" is general for dienes and trienes, and is even more pronounced for enone chromophores.

Absorption Intensity
Molar absoptivities may be very large for strongly absorbing chromophores (>10,000) and very small if absorption is weak (10 to 100). The magnitude of reflects both the size of the chromophore and the probability that light of a given wavelength will be absorbed when it strikes the chromophore. A general equation stating this relationship may be written as follows: = 0.87 1020 a ( where in cm2 ) is the transition probability ( 0 to 1 ) & a is the chromophore area

The factors that influence transition probabilities are complex, and are treated by what spectroscopists refer to as "Selection Rules". A rigorous discussion of selection rules is beyond

the scope of this text, but one obvious factor is the overlap of the orbitals involved in the electronic excitation. This is nicely illustrated by the two common transitions of an isolated carbonyl group. The n __> * transition is lower in energy ( max=290 nm) than the __> * transition ( max=180 nm), but the of the former is a thousand times smaller than the latter. The spatial distribution of these orbitals suggests why this is so. As illustrated in the following diagram, the n-orbitals do not overlap at all well with the * orbital, so the probability of this excitation is small. The __> * transition, on the other hand, involves orbitals that have significant overlap, and the probability is near 1.0.

Empirical Rules for Absorption Wavelengths of Conjugated Systems


Woodward-Fieser Rules for Calculating the Polyenes Core Chromophore
max

of Conjugated Dienes and

Substituent and Influence R- (Alkyl Group) .... +5 nm RO- (Alkoxy Group) .. +6 X- (Cl- or Br-) ......... +10 RCO2- (Acyl Group) .... 0 RS- (Sulfide Group) .. +30 R2N- (Amino Group) .. +60 Further -Conjugation C=C (Double Bond) ... +30 C6H5 (Phenyl Group) ... +60

Transoid Diene 215 nm

Cyclohexadiene* 260 nm

(i) Each exocyclic double bond adds 5 nm. In the example on the right, there are two exo-double bond components: one to ring A and the other to ring B. (ii) Solvent effects are minor. * When a homoannular (same ring) cyclohexadiene chromophore is present, a base value of 260 nm should be chosen. This includes the ring substituents. Rings of other size have a lesser influence. (calculated) = Base (215 or 260) + Substituent Contributions

max

Some examples that illustrate these rules follow.

Woodward-Fieser Rules for Calculating the __> Compounds Core Chromophore R = Alkyl 215 nm R = H 210 nm R = OR' 195 nm

max

of Conjugated Carbonyl

Substituent and Influence - Substituent R- (Alkyl Group) +10 nm Cl- (Chloro Group) +15 Br- (Chloro Group) +25 HO- (Hydroxyl Group) +35 RO- (Alkoxyl Group) +35 RCO2- (Acyl Group) +6 - Substituent R- (Alkyl Group) +12 nm Cl- (Chloro Group) +12 Br- (Chloro Group) +30 HO- (Hydroxyl Group) +30 RO- (Alkoxyl Group) +30 RCO2- (Acyl Group) +6 RS- (Sulfide Group) +85 R2N- (Amino Group) +95 & - Substituents R- (Alkyl Group) +18 nm (both & ) HO- (Hydroxyl Group) +50 nm ( ) RO- (Alkoxyl Group) +30 nm ( )

Cyclopentenone 202 nm

Further -Conjugation C=C (Double Bond) ... +30 C6H5 (Phenyl Group) ... +60 (i) Each exocyclic double bond adds 5 nm. In the example on the right, there are two exo-double bond components: one to ring A and the other to ring B. (ii) Homoannular cyclohexadiene component adds +35 nm (ring atoms must be counted separately as substituents) (iii) Solvent Correction: water = 8; methanol/ethanol = 0; ether = +7; hexane/cyclohexane = +11 (calculated) = Base + Substituent Contributions and Corrections

max

Some examples that illustrate these rules follow.

Das könnte Ihnen auch gefallen