Sunteți pe pagina 1din 27

1.1: What is Analytical Chemistry?

Analytical chemistry is too broad and too active a discipline for us to define completely. This
description is misleading. In this chapter we will try to say a little about what analytical
chemistry is, as well as a little about what analytical chemistry is not. Analytical chemistry is
often described as the area of chemistry responsible for characterizing the composition of matter,
both qualitatively (Is there any lead in this sample?) and quantitatively (How much lead is in this
sample?).
1.2: The Analytical Perspective
Many analytical chemists describe this perspective as an analytical approach to solving
problems. Although there are probably as many descriptions of the analytical approach as there
are analytical chemists, it is convenient for our purpose to define it as the five-step process.
1.3: Common Analytical Problems
This is the scope of a qualitative analysis is to identify what is present in a sample. Perhaps the
most common analytical problem is a quantitative analysis. The purpose of a qualitative,
quantitative, or characterization analysis is to solve a problem associated with a particular
sample. The purpose of a fundamental analysis, on the other hand, is to improve our
understanding of the theory behind an analytical method

Basic Tools of Analytical Chemistry


2.1: Measurements in Analytical Chemistry
Analytical chemistry is a quantitative science. Whether determining the concentration of a
species, evaluating an equilibrium constant, measuring a reaction rate, or drawing a correlation
between a compound’s structure and its reactivity, analytical chemists engage in “measuring
important chemical things.”
2.2: Concentration
Concentration is a general measurement unit stating the amount of solute present in a known
amount of solution
Concentration=amount of soluteamount of solution(2.1)(2.1)Concentration=amount of
soluteamount of solution
Although we associate the terms “solute” and “solution” with liquid samples, we can extend their
use to gas-phase and solid-phase samples as well. Table 2.4 lists the most common units of
concentration.
Table 2.4 Common Units for Reporting
Concentration

Name Units Symbol

Molessolute/literssolution
molarity M

formality Molessolute F
/ literssolution

normality N
Equivalentssolute /
literssolution

molality Molessolute m
/ kilogramssolvent

weight Gramssolute / % w/w


percent 100gramssolution

volume mLsolute / % v/v


percent 100mLsolution

weight- Grams solute % w/v


to- / 100mLsolution
volume
percent
parts per Grams ppm
million solute/10^6gramssolution

parts per Grams ppb


billion solute/10^9gramssolution

An alternative expression for weight percent is


Gramssolute/gramssolution×100
You can use similar alternative expressions for volume percent and for weight-to-volume
percent.
Gravimetric Methods
Gravimetric analysis is a technique through which the amount of an analyte (the ion being
analyzed) can be determined through the measurement of mass. Gravimetric analyses depend on
comparing the masses of two compounds containing the analyte. The principle behind
gravimetric analysis is that the mass of an ion in a pure compound can be determined and then
used to find the mass percent of the same ion in a known quantity of an impure compound. In
order for the analysis to be accurate, certain conditions must be met:
The ion being analyzed must be completely precipitated.
The precipitate must be a pure compound.
The precipitate must be easily filtered.
An example of a gravimetric analysis is the determination of chloride in a compound. In order to
do a gravimetric analysis, a cation must be found that forms an insoluble compound with
chloride. This compound must also be pure and easily filtered. The solubility rules indicate that
Ag+, Pb2+, and Hg22+ form insoluble chlorides. Therefore silver chloride could be used to
determine % Cl-, because it is insoluble (that is, about 99.9% of the silver is converted to AgCl)
and it can be formed pure and is easily filtered.
Put enough unknown into a weighing bottle with the lid on sideways (see Figure 1 below) and
dry in the oven. Cool in a desiccator.
Figure 1. Sample in a weighing bottle. Note the position of the lid for heating.
Indirectly weigh some mass, determined to 0.1 mg, of unknown into beaker.
Dissolve the unknown.
Add a precipiating agent to the solution (see Fig. 2 below).

Figure 2. The addition of a precipitating agent.

Figure 3. Heating the solution.


Optional - heat the solution on a hot plate to increase the particle size for easier filtering (see
Figure 3 above). This is usually referred to as digestion.
Test for complete precipitation by adding a drop of the precipitating agent and looking for any
sign of precipitate. Click for videos of complete and incomplete precipitation.
Filter the solution using vacuum filtration. Use a rubber policeman (click for video) to make sure
all the precipitate has been transferred from the beaker to the filter. It is important that the
precipitate is quantitatively transferred to the filter. If any remains in the beaker, the mass
obtained will be inaccurate.
Dry and weigh the precipitate.
Use stoichiometry to determine the mass of the ion being analyzed.
Find percent by mass of analyte by dividing the mass of the anayte by the mass of the unknown.
The following calculations would be done for the gravimetric determination of chloride:
Mass of sample of unknown chloride after drying: 0.0984 g
Mass of AgCl precipitate: 0.2290 g
One mole of AgCl contains one mole of Cl-. Therefore:
(0.2290 g AgCl) / (143.323 g/mol) = 1.598 x 10-3 mol AgCl
(1.598 x 10-3 mol AgCl) x (35.453 g/mol Cl) = 0.0566 g Cl
(0.0566 g Cl) / (0.0984 g sample) x 100% = 57.57% Cl in unknown chloride sample
Notice that even though the mass of sample (0.0984) only contains three significant figures, the
number is known to one part in a thousand (0.0001/0.0984 = 1/1000). The number 0.0984
therefore actually is "good" to four significant figures and the answer can be expressed to four
significant figures.
If Pb2+ had been used to precipitate the chloride, the calculuation would need to be modified to
account for the fact that each mole off PbCl2 contains two moles of chloride. Lead would not be
a good precipitating reagent, however, because PbCl2 is moderately soluble and therefore a
small amount of chloride would remain in solution, rather than in the precipitate.

Types of Gravimetric Analysis


There are three types of gravimetric analysis which are listed below:
Precipitation Gravimetric Analysis
Volatilization Gravimetric Analysis
Particulate Gravimetric Analysis

Precipitation Gravimetry
In precipitation gravimetry an insoluble compound forms when we add a precipitating reagent, or
precipitant, to a solution containing our analyte. In most methods the precipitate is the product of
a simple metathesis reaction between the analyte and the precipitant; however, any reaction
generating a precipitate can potentially serve as a gravimetric method.

Volatilization Gravimetry
A second approach to gravimetry is to thermally or chemically decompose the sample and
measure the resulting change in its mass. Alternatively, we can trap and weigh a volatile
decomposition product. Because the release of a volatile species is an essential part of these
methods, we classify them collectively as volatilization gravimetric methods of analysis.

Particulate Gravimetry
Precipitation and volatilization gravimetric methods require that the analyte, or some other
species in the sample, participate in a chemical reaction. In a direct precipitation gravimetric
analysis, for example, we convert a soluble analyte into an insoluble form that precipitates from
solution. In some situations, however, the analyte is already present as in a particulate form that
is easy to separate from its liquid, gas, or solid matrix.
Titrimetric Methods
Titrimetry, in which volume serves as the analytical signal, made its first appearance as an
analytical method in the early eighteenth century. Titrimetric methods were not well received by
the analytical chemists of that era because they could not duplicate the accuracy and precision of
a gravimetric analysis. Not surprisingly, few standard texts from the 1700s and 1800s include
titrimetric methods of analysis.
Precipitation gravimetry developed as an analytical method without a general theory of
precipitation. An empirical relationship between a precipitate’s mass and the mass of analyte—
what analytical chemists call a gravimetric factor—was determined experimentally by taking a
known mass of analyte through the procedure. Today, we recognize this as an early example of
an external standardization. Gravimetric factors were not calculated using the stoichiometry of a
precipitation reaction because chemical formulas and atomic weights were not yet available!
Unlike gravimetry, the development and acceptance of titrimetry required a deeper
understanding of stoichiometry, of thermodynamics, and of chemical equilibria. By the 1900s,
the accuracy and precision of titrimetric methods were comparable to that of gravimetric
methods, establishing titrimetry as an accepted analytical technique.

Acid-Base Titrations
Acid–base titrations, in which an acidic or basic titrant reacts with a titrand that is a base or an
acid, is probably the most common titration used by students in laboratories. To understand the
relationship between an acid–base titration’s end point and its equivalence point we must know
how the pH changes during a titration.
The Titration Experiment
Titration is a general class of experiment where a known property of one solution is used to infer
an unknown property of another solution. In acid-base chemistry, we often use titration to
determine the pH of a certain solution.
A setup for the titration of an acid with a base is shown in :

Figure %: A titration setup


We use this instrumentation to calculate the amount of unknown acid in the receiving flask by
measuring the amount of base, or titrant, it takes to neutralize the acid. There are two major ways
to know when the solution has been neutralized. The first uses a pH meter in the receiving flask
adding base slowly until the pH reads exactly 7. The second method uses an indicator. An
indicator is an acid or base whose conjugate acid or conjugate base has a color different from
that of the original compound. The color changes when the solution contains a 1:1 mixture of the
differently colored forms of the indicator. As you know from the Henderson-Hasselbalch
equation, the pH equals the pK a of the indicator at the endpoint of the indicator. Since we know
the pH of the solution and the volume of titrant added, we can then deduce how much base was
needed to neutralize the unknown sample.
Titration Curves
A titration curve is drawn by plotting data attained during a titration, titrant volume on the x-axis
and pH on the y-axis. The titration curve serves to profile the unknown solution. In the shape of
the curve lies much chemistry and an interesting summary of what we have learned so far about
acids and bases.
The titration of a strong acid with a strong base produces the following titration curve:

Figure %: Titration curve of a strong base titrating a strong acid


Note the sharp transition region near the equivalence point on the . Also remember that the
equivalence point for a strong acid-strong base titration curve is exactly 7 because the salt
produced does not undergo any hydrolysis reactions.
However, if a strong base is used to titrate a weak acid, the pH at the equivalence point will not
be 7. There is a lag in reaching the equivalence point, as some of the weak acid is converted to its
conjugate base. You should recognize the pair of a weak acid and its conjugate base as a buffer.
In , we see the resultant lag that precedes the equivalence point, called the buffering region. In
the buffering region, it takes a large amount of NaOH to produce a small change in the pH of the
receiving solution.

Figure %: Titration curve of a strong base titrating a weak acid


Because the conjugate base is basic, the pH will be greater than 7 at the equivalence point. You
will need to calculate the pH using the Henderson-Hasselbalch equation, and inputting the
pK b and concentration of the conjugate base of the weak acid.
The titration of a base with an acid produces a flipped-over version of the titration curve of an
acid with a base. pH is decreased upon addition of the acid.
Note that the pH of a solution at the equivalence point has nothing to do with the volume of
titrant necessary to reach the equivalence point; it is a property inherent to the composition of the
solution. The pH at the equivalence point is calculated in the same manner used to calculate the
pH of weak base solutions in Calculating pH's.
When polyprotic acids are titrated with strong bases, there are multiple equivalence points. The
titration curve of a polyprotic acid shows an equivalence point for the each protonation:

Figure %: Titration curve of a strong base titrating a polyprotic acid


The titration curve shown above is for a diprotic acid such as H2SO4 and is not unlike two
stacked . For a diprotic acid, there are two buffering regions and two equivalence points. This
proves the earlier assertion that polyprotic acids lose their protons in a stepwise manner.

Complexion titration
•Complexometric titration is a type of titration based on complex formation between the analyte
and titrant
• Complexometric titrations are particularly
Useful for determination of a mixture of different metal ions in solution. An indicator with a
marked color change is usually used to detect the end-point of the usually used to detect the
endpoint of the titration.
Any complexation reaction can in theory be applied as a volumetric technique provided that :
The reaction reaches equilibrium rapidly following each addition of titrant.
Interfering situations do not arise (such as stepwise formation of various complexes resulting in
the presence of more than one complex in solution in presence
Of more than one complex in solution in significant concentration during the titration process).
an complexometric indicator capable of locating equivalence point with fair accuracy is available
equivalence point with fair accuracy is available

Complexometric titration with EDTA


Ethylenediamminetetraacetic acid, has four carboxyl groups and two amine
groups that can act as electron pair donors or and two amine groups that can act as electron pair
donors, or Lewis bases.The ability of EDTA to potentially donate its six lone pairs of electrons
for the formation of coordinate covalent lone pairs of electrons for the formation of coordinate
covalent bonds to metal cations makes EDTA a hexadentate ligand. However in practice EDTA
is usually only partially ionized and
However, in practice EDTA is usually only partially ionized, and thus forms fewer than six
coordinate covalent bonds with metal cations Disodium EDTA commonly used in the
standardization cations. DisodiumEDTA, commonly used in the standardization of aqueous
solutions of transition metal cations, only forms four coordinate covalent bonds to metal cations
at pHm values less coordinate covalent bonds to metal cations at pH values less than or equal to
12 as in this range of pH values the amine groups remain protonated and thus unable to donate
electrons groups remain protonated and thus unable to donate electrons to the formation of
coordinate covalent bonds.
In analytical chemistry the shorthand "Na2H2Y" is typically used to designate disodium EDTA.
This shorthand can be used to designate disodium EDTA. This shorthand can be used to
designate any species of EDTA. The "Y" stands for the EDTA molecule, and the "Hn"
designates the number of acidic protons bonded to the EDTA molecule. EDTA forms an
octahedral complex with most 2+ metal cations M2+ in aqueous solution The main reason that
EDTA cations, M2+, in aqueous solution. The main reason that EDTA is used so extensively in
the standardization of metal cation solutions is that the formation constant for most metal cation-
EDTA complexes is very high, meaning that the equilibrium for the reaction :

2M2+ + H4Y → MH2Y + 2H+


lies far to the right. Carrying out the reaction in a basic buffer solution removes H+ as it is
formed which also drives the solution removes H as it is formed, which also drives the reaction
to the right. For most purposes it can be considered that the formation of the metal cation-EDTA
complex goes to
completion, and this is chiefly why EDTA is used in titrations / standardizations of this type.
To carry out met al cation titrations using EDTA it is almost always necessary to use a
complexometric indicator, usually an organic dye such as Fast Sulphon Black, Eriochrome Black
T, Eriochrome Red B or Murexide, to determine when the end point has been reached. These
dyes bind to the metal cations in solution to form colored complexes. However, since EDTA
binds to metal cations much more strongly than does the dye used as an indicator the EDTA will
displace the dye from the metal indicator the EDTA will displace the dye from the metal cations
as it is added to the solution of analyte. A color change in the solution being titrated indicates
that all of the dye has been displaced from the metal cations in
solution and that the endpoint has been reached solution, and that the endpoint has been reached.

•Complexometric titration has made it possible for man to be exposed to an advanced method of
titration which not only advanced method of titration which not only enables us to analyze more
ions, but also do them in very small quantities them in very small quantities.
• We’ve to be aware of the effects of pH on the titration method. Complex ion titration is
possible in very minute quantities The possible in very minute quantities. The biological use of
complexometric titration seems to involve an advanced method of seems to involve an advanced
method of this kind of titration; and we can learn its application on living cells.
The general shape of titration curves obtained by titrating 10.0 mL of a 0.01M solution of a
metal ion M with a 0.01M EDTA solution. The apparent stability constants of various metal-
EDTA complexes are indicated at the extreme right of the curves It is evident that the greater the
stability extreme right of the curves. It is evident that the greater the stability constant, the
sharper is the end point provided the pH is maintained constant.
Redox Titration
Titration is a very general way of using the reaction between two compounds to determine the
amount of one of them. You have already used titrations to determine the amount of acid present
in an unknown sample and,from this, the equivalent weight of the acid. Now you will use the
reaction between an oxidizing agent (KMnO4) and a reducing agent (Na2C2O4) to determine the
amount of sodium oxalate, Na2C2O4, in an unknown sample. This titration is particularly
convenient since it provides its own endpoint: as long as reducing agent is present, the purple
KMnO4 will be reduced to nearly colorless Mn2 +; when all of the reducing agent has been used,
the next purple drop will remain in solution to signal completion of the titration, that is, the
endpoint. To determine the number of equivalents of Na2C2O4 present in a sample, we need to
know the exact normality of the KMnO4 used in the titration, as well as the volume of the
oxidizing agent used to reach the endpoint. That is, we will need to standardize the KMnO4
solution first. The KMnO4( of unknown normality) is titrated against a known weight of pure
ferrous ammonium sulfate[Fe(NH4)2( S O4)2• 6H2O], a source of the Fe2 + ion. Once you
have determined the normality of the KMnO4 solution you can use it as the standard in your
titration of Na2C2O4 in your unknown. While the process of titration seems straightforward
enough, sources of error are common enough to make it a challenge (as you may have already
found in the acid-base titration experiment). Your first thoughts about sources of error probably
center on the accuracy of the endpoint. In practice, you are not likely to overshoot the endpoint
by more than a few drops, so we must look for other sources of error as well.

More likely chances for error include:


• Incomplete mixing of the KMnO4 solution initially. This will mean that each buret full will
have a different concentration.
• Weighing errors, simple misreading of the balance and/or loss of sample when transferring it to
the flask.
•Burette reading errors such as air bubbles in the tip, unrecorded initial readings, careless or
incorrect readings, and burette leakage during titration.
• Loss of sample in titration through splashing, incomplete dissolution of sample, or neglecting
to wash down the sides of the flask occasionally. Consistency of results, while no guarantee of
accuracy, is an encouraging sign.
Spectroscopic Methods
the branch of science concerned with the investigation and measurement of spectra produced
when matter interacts with or emits electromagnetic radiation.

Spectroscopy Based on Absorption


There are two general requirements for an analyte’s absorption of electromagnetic radiation.
First, there must be a mechanism by which the radiation’s electric field or magnetic field
interacts with the analyte. For ultraviolet and visible radiation, absorption of a photon changes
the energy of the analyte’s valence electrons. A bond’s vibrational energy is altered by the
absorption of infrared radiation. (Figure 10.3 provides a list of the types of atomic and molecular
transitions associated with different types of electromagnetic radiation.)
The second requirement is that the photon’s energy, hν, must exactly equal the difference in
energy, ∆E, between two of the analyte’s quantized energy states. Figure 10.4 shows a simplified
view of a photon’s absorption, which is useful because it emphasizes that the photon’s energy
must match the difference in energy between a lower-energy state and a higher-energy state.
What is missing, however, is information about what types of energy states are involved, which
transitions between energy states are likely to occur, and the appearance of the resulting
spectrum.
We can use the energy level diagram in Figure 10.15 to explain an absorbance spectrum. The
lines labeled E0 and E1 represent the analyte’s ground (lowest) electronic state and its first
electronic excited state. Superimposed on each electronic energy level is a series of lines
representing vibrational energy levels.

UV/Vis and IR Spectroscopy

The 1930s and 1940s saw the introduction of photoelectric transducers for ultraviolet and visible
radiation, and thermocouples for infrared radiation. As a result, modern instrumentation for
absorption spectroscopy became routinely available in the 1940s—progress has been rapid ever
since. Frequently an analyst must select—from among several instruments of different design—
the one instrument best suited for a particular analysis
Instrumentation
Frequently an analyst must select—from among several instruments of different design—the one
instrument best suited for a particular analysis. In this section we examine several different
instruments for molecular absorption spectroscopy, emphasizing their advantages and
limitations. Methods of sample introduction are also covered in this section.
Instrument Designs for Molecular UV/Vis Absorption
Filter Photometer. The simplest instrument for molecular UV/Vis absorption is a filter
photometer (Figure 10.25), which uses an absorption or interference filter to isolate a band of
radiation. The filter is placed between the source and the sample to prevent the sample from
decomposing when exposed to higher energy radiation. A filter photometer has a single optical
path between the source and detector, and is called a single-beam instrument. The instrument is
calibrated to 0% T while using a shutter to block the source radiation from the detector. After
opening the shutter, the instrument is calibrated to 100% T using an appropriate blank. The blank
is then replaced with the sample and its transmittance measured. Because the source’s incident
power and the sensitivity of the detector vary with wavelength, the photometer must be
recalibrated whenever the filter is changed. Photometers have the advantage of being relatively
inexpensive, rugged, and easy to maintain. Another advantage of a photometer is its portability,
making it easy to take into the field. Disadvantages of a photometer include the inability to
record an absorption spectrum and the source’s relatively large effective bandwidth, which limits
the calibration curve’s linearity.
Single-Beam Spectrophotometer.
An instrument that uses a monochromator for wavelength selection is called
a spectrophotometer. The simplest spectrophotometer is a single-beam instrument equipped with
a fixed-wavelength monochromator (Figure 10.26). Single-beam spectrophotometers are
calibrated and used in the same manner as a photometer. One example of a single-beam
spectrophotometer is Thermo Scientific’s Spectronic 20D+, which is shown in the photographic
insert to Figure 10.26. The Spectronic 20D+ has a range of 340–625 nm (950 nm when using a
red-sensitive detector), and a fixed effective bandwidth of 20 nm. Battery-operated, hand-held
single-beam spectrophotometers are available, which are easy to transport into the field. Other
single-beam spectrophotometers also are available with effective bandwidths of 2–8 nm. Fixed
wavelength single-beam spectrophotometers are not practical for recording spectra because
manually adjusting the wavelength and recalibrating the spectrophotometer is awkward and
time-consuming. The accuracy of a single-beam spectrophotometer is limited by the stability of
its source and detector over time.
Double-Beam Spectrophotometer.
The limitations of fixed-wavelength, single-beam spectrophotometers are minimized by using
a double-beam spectrophotometer (Figure 10.27). A chopper controls the radiation’s path,
alternating it between the sample, the blank, and a shutter. The signal processor uses the
chopper’s known speed of rotation to resolve the signal reaching the detector into the
transmission of the blank, P0, and the sample, PT. By including an opaque surface as a shutter, it
is possible to continuously adjust 0% T. The effective bandwidth of a double-beam
spectrophotometer is controlled by adjusting the monochromator’s entrance and exit slits.
Effective bandwidths of 0.2–3.0 nm are common. A scanning monochromator allows for the
automated recording of spectra. Double-beam instruments are more versatile than single-beam
instruments, being useful for both quantitative and qualitative analyses, but also are more
expensive.
Diode Array Spectrometer.
An instrument with a single detector can monitor only one wavelength at a time. If we replace a
single photomultiplier with many photodiodes, we can use the resulting array of detectors to
record an entire spectrum simultaneously in as little as 0.1 s. In a diode array spectrometer the
source radiation passes through the sample and is dispersed by a grating (Figure 10.28). The
photodiode array is situated at the grating’s focal plane, with each diode recording the radiant
power over a narrow range of wavelengths. Because we replace a full monochromator with just a
grating, a diode array spectrometer is small and compact.
Photoluminescence Spectroscopy

Photoluminescence is divided into two categories: fluorescence and phosphorescence. A pair of


electrons occupying the same electronic ground state have opposite spins and are said to be in a
singlet spin state (Figure 10.47a). When an analyte absorbs an ultraviolet or visible photon, one
of its valence electrons moves from the ground state to an excited state with a conservation of the
electron’s spin (Figure 10.47b). Emission of a photon from the singlet excited state to the singlet
ground state—or between any two energy levels with the same spin—is called fluorescence. The
probability of fluorescence is very high and the average lifetime of an electron in the excited
state is only 10–5–10–8 s. Fluorescence, therefore, decays rapidly once the source of excitation
is removed.
In some cases an electron in a singlet excited state is transformed to a triplet excited state (Figure
10.47c) in which its spin is no longer paired with the ground state. Emission between a triplet
excited state and a singlet ground state—or between any two energy levels that differ in their
respective spin states–is called phosphorescence. Because the average lifetime for
phosphorescence ranges from 10–4–104 s, phosphorescence may continue for some time after
removing the excitation source.
Fluorescence and Phosphorescence Spectra
To appreciate the origin of fluorescence and phosphorescence we must consider what happens to
a molecule following the absorption of a photon. Let’s assume that the molecule initially
occupies the lowest vibrational energy level of its electronic ground state, which is a singlet state
labeled S0 in Figure 10.48. Absorption of a photon excites the molecule to one of several
vibrational energy levels in the first excited electronic state, S1, or the second electronic excited
state, S2, both of which are singlet states. Relaxation to the ground state occurs by a number of
mechanisms, some involving the emission of photons and others occurring without emitting
photons. These relaxation mechanisms are shown in Figure 10.48. The most likely relaxation
pathway is the one with the shortest lifetime for the excited state.
?

Chromatography
a technique for the separation of a mixture by passing it in solution or suspension through a
medium in which the components move at different rates.
How it works
In all chromatography there is a mobile phase and a stationary phase. The stationary phase is
the phase that doesn't move and the mobile phase is the phase that does move. The mobile phase
moves through the stationary phase picking up the compounds to be tested. As the mobile phase
continues to travel through the stationary phase it takes the compounds with it. At different
points in the stationary phase the different components of the compound are going to be
absorbed and are going to stop moving with the mobile phase. This is how the results of any
chromatography are gotten, from the point at which
the different components of the compound stop moving and separate from the other components.
In paper and thin-layer chromatography the mobile phase is the solvent. The stationary phase in
paper chromatography is the strip or piece of paper that is placed in the solvent. In thin-layer
chromatography the stationary phase is the thin-layer cell. Both these kinds of chromatography
use capillary action to move the solvent through the
stationary phase.

What is the Retention Factor, Rf?


The retention factor, Rf, is a quantitative indication of how far a particular compound travels in a
particular solvent. The Rf value is a good indicator of whether an unknown compound and a
known compound are similar, if not identical. If the Rf value for the unknown compound is
close or the same as the Rf value for the known compound then the two compounds are most
likely similar or identical. The retention factor, Rf, is defined as
Rf= distance the solute (D1) moves divided by the distance traveled by the solvent front (D2)

Rf= D1/ D2
where

D1= distance that color traveled, measured from center of the band of color to the point where
the food color was applied

D2 = total distance that solvent traveled


The Different Types of Chromatography

There are four main types of chromatography. These are


Liquid Chromatography, Gas Chromatography, Thin Layer Chromatography and Paper
Chromatography.
Liquid Chromatography is used in the world to test water samples to look for pollution in lakes
and rivers. It is used to analyze metal ions and organic co mpounds in solutions. Liquid
chromatography uses liquids which may incorporate hydrophilic, insoluble molecules.

Gas Chromatography is used in airports to detect bombs and is used is forensics in many
different ways. It is used to analyze fibers on a persons body and also analyze blood found at a
crime scene. In gas chromategraphy helium is used to move a gaseous mixture through a column
of absorbent material.

Thin-layer Chromatography uses an absorbent material on flat glass or plastic plates. This is a
simple and rapid
method to check the purity of an organic compound. It is used to detect pesticide or insecticide
residues in food. Thin-layer chromatography is also used in forensics to analyze the dye
composition of fibers.

Paper Chromatography is one of the most common types of chromatography. It uses a strip of
paper as the stationary phase. Capillary action is used to pull the solvents up through the paper
and separate the solutes.

The table below summarizes the information from


above.

Type of Chromatography Applications in the Real Why and What is it


World
Liquid test water samples Used to analyze metal
Chromatography to look for ions and organic
pollution, compounds in solutions.
It uses liquids which
may incorporate
hydrophilic, insoluble
molecules.

Gas detect bombs in Used to analyze volatile


Chromatography airports, identify gases. Helium is used to
and quantify such move the gaseous
drugs as alcohol, mixture through a
used in forensics to column of absorbent
compare fibers material.
found on a victim

Thin-Layer detecting pesticide Uses an absorbent


Chromatography or insecticide material on flat glass
residues in food, plates. This is a simple
also used in and rapid method to
forensics to analyze check the purity of the
the dye organic compound.
composition of
fibers

Paper separating amino The most common type


Chromatography acids and anions, of chromatography. The
RNA paper is the stationary
fingerprinting, phase. This uses
separating and capillary action to pull
testing histamines, the solutes up through
antibiotics the paper and separate
the solutes.

Kinetic Methods
Another way to classify analytical techniques is by whether the analyte’s concentration is
determined by an equilibrium reaction or by the kinetics of a chemical reaction or a physical
process.
Kinetic Methods Versus Equilibrium Methods
In an equilibrium method the analytical signal is determined by an equilibrium reaction
involving the analyte or by a steady-state process that maintains the analyte’s concentration. In a
kinetic method the analytical signal is determined by the rate of a reaction involving the analyte,
or by a nonsteady-state process. As a result, the analyte’s concentration changes during the time
in which we are monitoring the signal.
Chemical Kinetics
Every chemical reaction occurs at a finite rate, making it a potential candidate for a chemical
kinetic method of analysis. To be effective, however, the chemical reaction must meet three
necessary conditions: the reaction must not occur too quickly or too slowly; we must know the
reaction’s rate law; and we must be able to monitor the change in concentration for at least one
species.
Reaction Rate
The rate of the chemical reaction—how quickly the concentrations of reactants and products
change during the reaction—must be fast enough that we can complete the analysis in a
reasonable time, but also slow enough that the reaction does not reach equilibrium while the
reagents are mixing. As a practical limit, it is not easy to study a reaction that reaches
equilibrium within several seconds without the aid of special equipment for rapidly mixing the
reactants.
Rate Law
The second requirement is that we must know the reaction’s rate law—the mathematical
equation describing how the concentrations of reagents affect the rate—for the period in which
we are making measurements. For example, the rate law for a reaction that is first order in the
concentration of an analyte, A, is
rate=−d[A]/dt=k[A]

where k is the reaction’s rate constant.

Classifying Chemical Kinetic Methods


Figure 13.2 provides one useful scheme for classifying chemical kinetic methods of analysis.
Methods are divided into two main categories: direct-computation methods and curve-fitting
methods. In a direct-computation method we calculate the analyte’s initial concentration, [A]0,
using the appropriate rate law. For example, if the reaction is first-order in analyte, we can use
Equation 13.2.213.2.2 to determine [A]0 if we have values for k, t, and [A]t. With a curve-fitting
method, we use regression to find the best fit between the data—for example, [A]t as a function
of time—and the known mathematical model for the rate law. If the reaction is first-order in
analyte, then we fit Equation 13.2.213.2.2 to the data using k and [A]0 as adjustable parameters.
Direct-Computation Fixed-Time Integral Methods
A direct-computation integral method uses the integrated form of the rate law. In a one-point
fixed-time integral method, for example, we determine the analyte’s concentration at a single
time and calculate the analyte’s initial concentration, [A]0, using the appropriate integrated rate
law. To determine the reaction’s rate constant, k, we run a separate experiment using a standard
solution of analyte. Alternatively, we can determine the analyte’s initial concentration by
measuring [A]t for several standards containing known concentrations of analyte and
constructing a calibration curve.

Direct-Computation Variable-Time Integral Methods


In a variable-time integral method we measure the total time, ∆t, needed to effect a specific
change in concentration for one species in the chemical reaction. One important application is for
the quantitative analysis of catalysts, which takes advantage of the catalyst’s ability to increase
the rate of reaction. As the concentration of catalyst increase, ∆t decreases. For many catalytic
systems the relationship between ∆t and the catalyst’s concentration is
1/Δt=Fcat[A]0+Funcat

where [A]0 is the catalyst’s concentration, and FcatFcat and FuncatFuncat are constants
accounting for the rate of the catalyzed and uncatalyzed reactions.

Curve-Fitting Methods
In direct-computation methods we determine the analyte’s concentration by solving the
appropriate rate equation at one or two discrete times. The relationship between the analyte’s
concentration and the measured response is a function of the rate constant, which we determine
in a separate experiment using a single external standard (see Example 13.1 or Example 13.2), or
a calibration curve (see Example 13.3 or Example 13.4).
In a curve-fitting method we continuously monitor the concentration of a reactant or a product as
a function of time and use a regression analysis to fit the data to an appropriate differential rate
law or integrated rate law. For example, if we are monitoring the concentration of a product for a
reaction that is pseudo-first-order in the analyte, then we can fit the data to the following
rearranged form of Equation 13.2.1913.2.19
[P]t=[A]0(1−e−k't)
using [A]0 and k′ as adjustable parameters. Because we are using data from more than one or
two discrete times, a curve-fitting method is capable of producing more reliable results.

S-ar putea să vă placă și