You are on page 1of 100

Polymer Analysis

Chapter 1. Introduction/Overview

The focus of this course is analysis and characterization of polymers and plastics. Analysis of polymeric systems is essentially a subtopic of the field of chemical analysis of organic materials. Because of this, spectroscopic techniques commonly used by organic chemists are at the heart of Polymer Analysis, e.g. infra-red (IR) spectroscopy, Raman spectroscopy, nuclear magnetic resonance (NMR) spectroscopy and to some extent ultra-violet/visible (UV/Vis) spectroscopy. In addition, since most polymeric materials are used in the solid state, traditional characterization techniques aimed at the solid state are often encountered, x-ray diffraction, optical and electron microscopy as well as thermal analysis. Unique to polymeric materials are analytic techniques which focus on viscoelastic properties, specifically, dynamic mechanical testing. Additionally, techniques aimed at determination of colloidal scale structure such as chain structure and molecular weight for high molecular weight materials are somewhat unique to polymeric materials, i.e. gel permeation chromatography, small angle scattering (SAS) and various other techniques for the determination of colloidal scale structure. The textbook, Polymer Characterization covers all of these analytic techniques and can serve as a reference for a general introduction to the analysis of polymeric systems. Due to time constraints we can only cover a small number of analytic techniques important to polymers in this course and these are outlined in the syllabus. Structure/Processing/Property: Generally people resort to analytic techniques when confronted with a problem which involves understanding the relationship between properties of a processed material and the structure and chemical composition of the system. Plastics are typically complex morphological systems being composed of many phases and additives, even the polymer itself being disperse in molecular weight, tacticity, crystallinity, orientation and sometimes chemical composition. Dispersion of structure and chemical composition means that the best tools to describe polymeric materials are always statistical in nature. For example, a low molecular weight organic has a specific melting point, while a polymer displays a range of melting with an onset, a peak and a maximum melting temperature. Such a melting spectrum might best be described by a Gaussian function with a standard deviation and mean. Additionally, the complication of enormous macromolecular chains means that simplified descriptions are often needed to characterize polymeric materials, for example group contribution methods in spectroscopy where the chain structure is decomposed into chemical groups which contribute to the absorption spectrum. A detailed understanding of many of the analytic descriptions of polymeric materials is often precluded by the complexity of the situation. Analytic Techniques: All analytic techniques used in polymer characterization are based on specific physical principles which serve as a guide to understanding the basic limitations of the techniques. Often multiple techniques are available to describe the same property of a material and it is only through understanding the physical basis of characterization techniques that one can pick the "right tool for the job", the job being understanding the relationship between structure and chemical composition and properties. For example, it is often found that blown films of polyethylene display different tear strengths in the machine and transverse directions. This can lead to failure of parts made from blown films such as plastic bags. It is often assumed that this directional nature of the tear strength is related to orientation induced by processing of these plastics. There are several analytic techniques available to describe orientation. These include construction of polefigures from x-ray diffraction scans, calculation of orientation functions from XRD data, calculation of orientation functions from IR, NMR or Raman data, and measurement of the optical
1

birefringence for the blown films using polarized light. Each of these techniques will yield a different value for the orientation function! Similarly, even the value for the degree of crystallinity in these blown films will depend on the technique which is used, i.e. XRD, differential scanning calorimetry, density, or IR for example. This makes a firm understanding of the physical basis of analytic techniques critical to their application in polymeric systems. Levels of Structure: The analytic description of a complex material is strongly dependent on the size scale on which an observation is made. For example, a semi-crystalline polymer such as polyethylene is composed of chemical units similar to olefinic waxes. These chemical units give rise to spectroscopic absorption patterns which are largely indistinguishable from their lower molecular weight counterparts. Similarly, the crystalline structure, which is usually of low symmetry in polymers, mimics the crystalline structure seen in lower molecular weight analogues such as waxes. The monomer structure combined with the topological arrangement of monomers in a polymer chain give rise to helical coiling of polymer chains. This helical coiling and the weak chemical associations related to it give rise to some mechanical and vibrational features which can often be observed spectroscopically. Generally, the helical coiling of monomers in a chain are evidenced by colloidal scale structure, chain persistence (local linearity) and enhancement of the ability of long chain polymers to crystallize. NMR is a technique particularly suited for the characterization of the topological arrangement of monomers in a chain (i.e. tacticity). In polyethylene, local chain structure is sufficiently regular to give rise to a crystalline phase. Entanglement of chains, chain branching and the presence of endgroups prevents complete crystallization of a polymer. Polymeric materials which display crystallinity are always described by a multi-phase model, i.e. semicrystalline, which includes an amorphous and crystalline phase in coexistence. Low transport coefficients and chain folding give rise to nano-scale crystallites which are best observed by TEM, small-angle x-ray scattering (SAXS) and to some extent by Raman spectroscopy (Longitudinal Acoustic Modes, LAM). Fibrillar crystallites in polymers lead to colloidal to optical scale structures, spherulites, which are generally centro-symetric and which display radially oriented birefringence. These micron scale structures are best observed using optical microscopy, SEM, small-angle light scattering (SALS) and give rise to certain features in the mechanical and transport properties of semi-crystalline polymers. Thus, is one in interested in a specific analytic feature of a processing operation such as film blowing on a polymeric material such as high density polyethylene (HDPE), one is immediately faced with the issue of structural level, i.e. for transport properties one might need to characterize the orientation of chains or crystallites (lamellae), for mechanical properties, orientation of spherulites and the amorphous component of these biphase materials. Additionally, it is expected that any analytic determination of these materials will be subject to a large range of statistical variation between samples as well as an innate distribution associated with the polydispersity of the material of itself. Course Contents: Statistics: All properties of polymeric systems display dispersion due to 1) the limited ability of synthetic chemistry to produce monodisperse and perfect chemical structure as well as 2) the dominance of
2

kinetics in processing of high molecular weight materials. Statistical analysis is critical to understanding and describing plastics. This course begins with a survey of the important tools to describe statistically distributed systems which includes the major distribution functions and mathematical descriptions of the propagation of error in data sets. Analytic descriptions of polymers are of no use unless some description of the expected error associated with the analytic results are presented. Thermal and Mechanical Testing: Thermal analysis is useful in describing solid state transitions in polymers and is of pivotal importance to understanding mechanical properties and processing of plastics. The measurement of the glass transition and crystalline transitions using typical analytic techniques will be discussed. Polymeric systems are dominated by kinetics and this is emphasized in the use of dynamic mechanical, thermal analysis techniques to describe mechanical absorption phenomena associated with the glass transition and other mechanical transitions. Absorption Spectroscopy: The major techniques for the determination of chemical composition and molecular topology involve the absorption of electro-magnetic radiation by polymers. The major techniques are IR, Raman, and NMR spectroscopy and the bulk of this course will involve these major analytic techniques. Absorption is a quantized inelastic phenomena involving the transfer of energy from EM radiation to a material. Diffraction and Scattering: In addition to inelastic absorption phenomena, elastic interaction between EM radiation and a material is possible and this gives rise to diffraction and scattering phenomena. The small crystallite size and dominance of crystalline orientation in processed plastics lead to several unique analytic approaches in the analysis of x-ray diffraction data in these materials. The focus in this course will be on those tools used in diffraction which are specific to polymeric materials. Small-angle x-ray scattering is a critical technique for the description of polymeric materials since diffraction at small angles is associated with the colloidal to nano-scales which is the size range of a typical polymer chain. The colloidal scale is also associated with polymer crystallites (lamellae) and microphase separated block copolymer structures. Further, light scattering has been widely used in polymer science to describe disordered micron scale structure as well as a primary technique for the determination of molecular weight from dilute solutions.

Polymer Analysis
Chapter 2. Error Analysis/Statistical Descriptions of Data.

All Polymer Properties are Disperse: Polymeric materials are subject to dispersion in all analytic properties. For example, the melting point in a low molecular weight organic or inorganic is a fixed value constant which might display some variability over 1 or 2 degrees. The melting point for high-density polyethylene (HDPE), for example can vary from about 110C to about 160C depending on processing and generally displays a broad dispersion over about 10 to 20C. Spectroscopic analysis of polymers using techniques such as infra red adsorption (IR) and nuclear magnetic resonance (NMR) rely mostly on local chemical groups, so can display fairly sharp absorption bands. In these spectroscopic techniques, dispersion is shown by the existence of peak splittings or the presence of a number of chemical species in small amounts. Additionally, many absorption bands in polymers are difficult to describe analytically and pertain to various acoustic modes associated with the conformation of long chain structures. These are typically broad bands due to wide dispersion in these long chain conformations. X-ray diffraction from polymer crystallites generally displays broad diffraction peaks associated with small crystals and a high degree of disorder within crystallites. Polymers also display dispersion in structural and chemical orientation which should be viewed with statistics. Dispersions in mechanical properties are always seen in polymers. Statistics are also used to describe the dispersion of chain size, molecular weight, and topological arrangement of tacticity. All descriptions of polymer chain size are based on statistics. It is critical to realized at the onset of any analysis of a polymeric system that every physical property will be described by a distribution. Error Analysis in Analytic Methods: In the physical sciences each analytic measurement must be associated with an assessment of the confidence which should be associated with the analytic description. A value for some property of a material is of no use unless some estimate of the expected error and distribution is provided. For example, for a commercial sample of HDPE a technician might report the melting point of a blown film as 135C. You might be involved in using this material in a packaging application where the material will be subjected to shipping at a maximum temperature of 95C. From the reported melting point and operating temperature would you feel confident that this material would meet the specs of the packaging company? The correct answer is that the measured value is of no use without a description of the statistical distribution in the samples as well as a statistical description of the distribution of melting points, i.e. onset, peak and maximum melting point. The technician should be required to measure at least 10 samples (preferably more) and calculate a standard deviation for the reported value. Additionally, similar determinations of the onset of melting and the maximum melting point would be needed to determine if this material meets the specifications of the packager. There are a number of useful texts which describe the correct handling of experimental data. The most commonly cited reference is "Data Reduction and Error Analysis for the Physical Sciences" by P. R. Bevington, 1969 McGraw Hill. This section of the course will summarize some of the major points of Bevington with an emphasis on applications in polymeric systems. Types of Error: Statisticians categorize three types of error. This framework may be useful in considering a specific experiment such as determination of the d-spacing from an XRD peak. Illegitimate Error: These errors involve an operator error, e. g. you have placed the wrong sample in the diffractometer and are determining the d-spacing for the wrong sample. There is no
1

statistical description for these errors unless you want to consider sociology. Your main protection against illegitimate errors is to always consider, when faced with extremely unexpected results, that the results involve a human error. Always look carefully at completely unexpected results. Systemic Error: There errors are also not subject to a statistical analysis. They result from faulty calibration of the instrument or other problems which result in a constant shift of the data. In XRD a systematic error might involve confusion of 2 with leading to roughly a doubling of the dspacings. Systemic errors can sometimes be corrected after the fact if one is careful. It is important to keep a good record of the analysis that was performed partly for the purpose of correction of systemic errors of this type. Random Error: Random errors are the main area which statistics can deal with (i.e. standard deviation, mean). Random errors have two sources. 1) Random variability in the samples themselves. 2) Limited precision of the analytic equipment. These two sources can be distinguished if more precise equipment is available or a standard sample is available. In describing the error involved in an analytic measurement all types of error should be considered. Accuracy of a measurement pertains to how close a measurement is to the actual value. Precision refers to the reproducibility of the measurement whether or not it is close to the actual value. If systematic error is present a measurement might be extremely precise but completely inaccurate! Also, if the equipment is well calibrated a measurement could be extremely accurate but not particularly precise. Mostly the importance of these terms is in communication of your confidence in a particular value and what the possible sources of error are. Statistical Analysis of Measurements: Generally a given measurement will be conducted several times in order to determine the statistical distribution for the measurement. The values most commonly determined are the mean, , and the standard deviation, :

1 N xi N 1 1 N ( xi ) 2 N 1 1

2 =

N is the number of measurements made, the value from each measurement is xi for measurement i. The square of the standard deviation is called the variance. Distribution Functions (see P. C. Heimenz, Polymer Chemistry, 1984, Marcel Dekker, pp. 34): If a measurement is not single valued then it is common to fit the number distribution of values to a distribution function. The simplest distribution functions will involve two parameters, the mean, and the standard deviation, . More complicated continuous distribution functions will involve higher moments of the distribution. The k'th moment of a distribution is given by: k' th moment =

f (x
i i

xs )

where fi is the fraction of all measurements which have the value x i, i.e. N i/N, xs is basis for the moment and k is the order of the moment. For example, the mean, , is the first moment (k = 1) about the origin (xs = 0). The variance, 2, is the second moment (k = 2) about the mean (xs = ). The molecular weight distributions commonly used in polymer science can be described in terms of the more broadly used statistical description of moments. The number average molecular weight, Mn is the same as the first moment about the origin or the mean. The weight average molecular weight (mass average), Mw, is given by:

Mw

N M = N M
i i i i

2 i

fM = fM
i i i i

2 i

or the ratio of the second to the first moment about the origin. The polydispersity index, Mw/Mn, is the ratio of the second moment to the square of the first moment about the origin. Given Mw and Mn the standard deviation of a distribution of molecular weights can be determined: M = Mn w 1 Mn
1 2

Several specific distribution functions are mentioned below. A distribution is usually considered unimodal if one of these continuous distribution functions describes the distribution of measured values. If the data is best described by several of these functions it is termed bimodal, trimodal etc. A bimodal distribution in lamellar thicknesses in polyethylene might be generated if crystallization occurred in two distinct steps such as primary crystallization and secondary epitaxial spherulitic decoration for instance. This distribution might be described by two Gaussian functions, so a total of 4 parameters, 2 means and 2 standard deviations. Binomial Distribution: Consider a tactic polymer such as polypropylene. As discussed in class, on passing along the main chain of the polymer there is a choice in handedness for the substitutent methyl group which could be considered similar to a coin toss, i.e. if the substitutent occurs with the same handedness as the previous mer unit this would be similar to a coin tossed heads (R for tacticity) and if it occurs with the opposite handedness (S for tacticity) this would be similar to a coin tossed tails. If the tacticity of the polymer is determined at random (atactic) then there is equal probability for R and S handedness. The binomial distribution describes such a situation where the "probability of success" is given by p, (consider R a success, here p = 0.5 for atactic polymers). The probability of observing "n" R handedness mer units in a chain of N mer units when the probability of seeing an R handiness mer unit is p is given by: PBinomial ( N , n, p) = N! N n p n (1 p) n!( N n)!
3

the mean, , and standard deviation, , for the binomial distribution are given by:

= Np 2 = Np(1 p)
PBinomial could be used to plot a distribution function of tacticities for example. Note that there is a finite probability for a completely R polymer (0! =1). Poisson Distribution: When N is large and the mean, , is constant with sample size, the binomial distribution simplifies to the Poisson Distribution: PPoisson (n, ) =

n e n!

where the standard deviation is given by:

=
The Poisson Distribution is used to describe small samples of large populations, so is appropriate in analytic techniques which involve counting of events, such as XRD and light counters, and mass spectrometers which measure events. The standard deviation in such a counting measurement (counting of events) is the square root of the number of counts. For example, an integrated diffraction peak might generate 10,000 counts on a proportional detector. The error in this value is 100. If an analytic instrument does not report "counts" or "events" then the Poission distribution value for the standard deviation can not be used directly. Gaussian Distribution: For very large samples, N, with a finite probability of success, p, a smooth distribution is usually observed. Such a distribution can be described by a Gaussian function: 1 n 2 1 PGaussian (n, , ) = exp 2 2 The Gaussian distribution is the basis of the "bell-shaped curve". It is also used to describe random walks in diffusion as well as the path of a polymer coil under theta-conditions. Integration of the Gaussian distribution as a weighting factor for r2 , the square of the chains end-to-end distance yields the mean square end-to-end distance for a "Gaussian" chain, Nl 2, where l is the step size and N is the number of randomly arranged steps in the non-interacting chain. For the Gaussian chain = 0 and 2 = nl2. Other Distributions: There are many other distribution functions commonly used in polymer science such as the Lorentzian Distribution (see hand out) and the Maxwellian Distribution (see Polymer Materials Science by J. Schultz for instance). The definition of these distributions will rely on an understanding of the moments of a continuous distribution described above.
4

Covariance: In many analytic experiments several parameters are determined from a single measurement through the use of a fit to experimental data. For instance, a light scattering curve can be used to determine the molecular weight (first moment about the origin) and second viral coefficient, A 2, through the Zimm plot (figure 2.8 in our text). Often the two parameters which are measured are not completely independent in terms of the fit to the data. Under such conditions it is necessary to determine the covariance of the two parameters. The covariance reflects the degree to which two parameters effect each other. If more than two parameters are unknown, then a covariance matrix can be constructed to determine the extent two which any two parameters are associated. For two parameters the covariance is defined by:
2 uv = (u u

)((v v ))

where <> indicates a mean. Covariance is included for completeness here. Determination of the covariance is generally rare in the literature. Propagation of Errors: In many analytic techniques a value is measured, an error is determined, and this value is used to calculate a parameter of interest. For example, in the determination of the modulus of a sample the extension of the sample is measured with some error and is normalized against the original length to determine the engineering strain. The force applied to the sample is measured with an experimental error and is normalized by the cross sectional area to determine the stress. These two parameters are plotted (stress versus strain) and a curve fit is used to determine the modulus at low strain. In order to determine the standard deviation in the modulus the experimentally determined errors in length, force and area must be propagated to the stress and strain and the error in these parameters must be propagated to a linear fit of the data points. Propagation of errors is a rudimentary tool necessary to perform polymer analysis. Consider a measurement of the absorption, A, and absorption coefficient, , using a single wavelength of light which passes through a sample and a photomultiplier tube which reports counts. The relevant equation is the Beer-Lambert Law for linear absorption (pp. 40 and 54 in Cambell): 1 I A = ln 0 = ln = cl I T where c is the concentration of absorbing species and l is the sample thickness. T is called the transmission. For a solid sample c = 1. Two measurements are necessary, I0 with no sample and I with a sample of thickness l. If the two measurements are I0 = 100,000 counts , and I = 20,000 counts, the respective standard deviation is given by a Poission Distribution function as the square root of the number of events, I0 = 100,000 320 counts; I = 20,000 140 counts. To propagate the error in I0 and I to 1/T the general error propagation rule in differential form can be used:

x 2 x 2 x x = + v + 2 uv u v u v
2 2 2 x 2 u

Here it is safe to assume that I and I0 are uncorrelated (unrelated) so the covariance, uv = 0. The two standard deviations are given above. For x = 1/T, x/I0 = 1/I, and x/I = -I0/I2. The above equation yields:

2 1/ T

I20
I2

2 2 I0 + 4I I

or

I20 I2 12/ T = + (1 / T )2 I02 I 2


this follows the general rule for ratios given by Bevington (see handout). The propagated value and error for 1/T is, 1/T = 5 0.04. The absorption, A, is the natural log of 1/T and the propagated error is given by:

=
2 A

2 1/ T

12/ T A = (1 / T ) (1 / T )2
2

So, A = 1.61 0.01. The sample thickness, l, is 0.1 0.02 cm as calculated from a series of 10 measurements using the mean and number of samples equation given above. The absorption coefficient, , is calculated from = A/cl, where c is 1 for a solid sample. Replacing the variables in the equation above for the error in 1/T, = 16 3 . Notice the reduction in the number of significant figures associated with the large error. It should also be noted that often the largest source of error is related to factors which are of minor significance to the measurement, e.g. here error in the sample thickness dominates the error in the absorption coefficient. Purpose of Error Analysis: Although it is assumed that you can determine the error value for any quantity you measure in science, the most important part of error analysis involves interpretation of the significance of the analysis in view of the error and in view of logic and the reasonableness of the results. Error analysis is a critical factor in both demonstrating the scientific reasonableness of a result, as well as forming a basis of a scientific critique of work performed for you by a technician or outside lab. Error analysis is always at the heart of a scientific argument. A measured value has no meaning without an analysis of the associated error. Such an analysis can be quantitative, as above, or qualitative, based on your scientific judgment. The value in either case is based on the reasonableness and logic of the approach. In the remainder of this course special emphasis will be given to qualitative and quantitative assessments of the error in the analytic techniques covered.
6

Least-Squares Fits: In order to determine analytic parameters it is often necessary to perform a curve fit to a raw data set. For example the determination of modulus from a stress-strain measurement requires a linear fit of the type = E where s is the stress, = F/A, and e is the strain, = l/l, and E is the Youngs modulus. The error in the stress measurement could be propagated from the estimated error in the Force and area measurements. These errors can be propagated to the modulus in a linear fit through a least squares minimization of 2. 2 is a measure of the difference between the actual data points and the projected points associated with the fit parameter. It bears resemblance to the variance,

=
2 1

( yi f ( xi ))2
i2

The propagated uncertainty in the coefficients for the least squares fit can be obtained in a computer program by calculation of the second derivative of 2 with respect to variation in the parameter:
2 = a j

2 a2 j
2 2

A full discussion of least-squares fitting routines and propagation of error is given in Bevington. For linear functions a relatively simple algorithm for propagation of error is given in the handout.

Chapter 3. Thermal Analysis (Chapter 12 Campbell & White). Polymers typically display broad melting endotherms and glass transitions as major analytic features associated with their properties. Both the glass and melting transitions are strongly dependent on processing conditions and dispersion in structural and chemical properties of plastics. Characterization of polymers requires a detailed analysis of these characteristic thermal transitions using either differential scanning calorimeter (DSC) or differential thermal analysis (DTA). Additionally, polymers are viscoelastic materials with strong time and temperature dependencies to their mechanical properties. Temperature scans across the dynamic spectrum of mechanical absorptions are commonly required for characterization of polymers, especially for elastomers. These thermal/mechanical properties are characterized in dynamic mechanical/thermal analysis (DMTA). Additionally, weight loss with heating is a common phenomena for polymers due to degradation and loss of residual solvents and monomers. Weight loss on heating is studied using thermal gravimetric analysis (TGA). A complete thermal analysis of a plastic sample yields inferential information concerning the chemical composition and structure of the material. Examples: 1. The Hoffman-Lauritzen description of the crystalline melting point associates shits in the melting transition temperature with the thickness of lamellar crystallites in polymers. Such structural based shifts would suggest further study using the Scherrer approach for diffraction peak broadening and small-angle x-ray and TEM analysis. 2. Dramatic weight loss in a TGA analysis of nylon at temperatures above 100C indicate some association of water with the nylon chemical structure. Such an observation would suggest further study using spectroscopic techniques. 3. In a polymer alloy (blend), the observation of two glass transition temperatures indicates a biphasic system, a single glass transition, a miscible system following the Flory-Fox equation. Further support for miscibility would come from microscopy and scattering (neutron, x-ray and light can all be used to characterize miscibility). Generally, thermal analysis is the easiest and most available of techniques to apply to a sample and for this reason thermal analysis is often the first technique used to analytically describe a plastic material. Error analysis in thermal techniques usually is conducted by repetition of the measurement for at least 3 to 10 identical samples in order to determine the standard deviation in the measurement. In all thermal analysis techniques the instruments must be calibrated with standard samples displaying sharp and constant transition temperatures and enthalpies of transition. Calorimetry (Differential Analysis, DTA): Scanning Calorimetry, DSC; Differential Thermal

Calorimetry involves the measurement of relative changes in temperature and heat or energy either under isothermal or adiabatic conditions. Chemical calorimetry where the heats of reaction are measured, usually involve isothermal conditions. Bomb or flame calorimeters involve adiabatic systems where the change in temperature can be translated, using the heat capacity of the system, into the enthalpy or energy content of a material such as in determination of the calorie content of food. In materials characterization calorimetry usually involves an adiabatic measurement. A calormetric measurement in materials science is carried out on a closed system where determination of the heat, Q, associated with a change in temperature, T, yields the heat capacity of the material, C:
1

C=

Q T

At constant pressure: dQ H = = CP dT T P, N The enthalpy can be calculated from CP through, H (T ) = H (T0 ) + CP dT


T0 T

Instrumentation for Thermal Analysis, DTA and DSC: Figure 12.1 of Campbell and White shows a schematic of a differential thermal analysis (DTA) instrument. The instrument is composed of two identical cells in which the sample and a reference (often an empty pan) are placed. Both cells are heated with a constant heat flux, Q, using a single heater, and the temperatures of the two cells are measured as a function of time. If the sample undergoes a thermal transition such as melting or glass transition a difference in temperature is observed, T = Tsample - Treference. Negative T indicates an endotherm for a heating cycle. Quantitative analysis of DTA data is complicated and the instrument is usually viewed as a fairly crude sibling of a differential scanning calorimeter (DSC) discussed below. Recent instrumental advancements have improved the quantitative use of DTA instruments. A DTA instrument is generally less expensive than a DSC. Determination of transition temperatures are accurate in a DTA. Estimates of enthalpies of transition are generally not accurate. In the DTA heat is provided at a constant rate and temperature is a dependent parameter. In the equation above for C P, the normal order of dependent and independent parameters in the differential is reversed, so dT/dQ is actually measured rather than dQ/dT. This distinction is critically important in transitions where kinetics become important such as in polymer melting and glass transition. Figure 12.2 of Campbell and White shows a schematic of a differential scanning calorimeter (DSC). The arrangement is similar to the DTA except that the sample and reference are provided with separate heaters. The independent parameter is the temperature which is ramped at a controlled rate. Feedback loops control the feed of heat to the sample and reference so the temperature program is closely followed. The raw data from a DSC is heat flux per time or power as a function of temperature at a fixed rate of change of temperature (typically 10C/min). Since the heat flux will increase with temperature ramp rate, higher heating rates lead to more sensitive thermal spectra. On the other hand, high heating rates lead to lower resolution of the temperature of transition and can have consequences for transitions which display kinetic features. Campbell and White go through a useful 2 page comparison of the DSC and DTA techniques which should be reviewed. Data Interpretation: The output of a DSC is a plot of heat flux (rate) versus temperature at a specified temperature ramp rate. The heat flux can be converted to CP by dividing by the constant rate of temperature change. The output from a DTA is temperature difference (T) between the reference and sample cells versus sample temperature at a specified heat flux. Qualitatively the two plots appear similar.
2

Both DSC's and DTA's must be calibrated, essentially, for each use since small changes in the sample cells (oil from fingers etc.) can significantly shift the instrumental calibration. For polymer samples these instruments are typically calibrated with low melting metal crystals that display a sharp melting transition such as indium (Tm = 155.8C) and low molecular weight organic crystals such as naphthalene. The volatility of low molecular weight organic crystals requires the use of special sealed sample holders. An instrumental time lag is always associated with scanning thermal analysis. The observed transitions may be "smeared" by this instrumental time lag (typically close to 1C at 10/min heating rate). Some account can be made for this time lag by comparison of results from different heating rates. Often this time lag is accounted for by taking the onset of melting as the melting point rather than the peak value for sharp melting standard samples used in calibration. For polymer samples, significant broadening of the melting peak (up to 25 to 50C) is the norm and this is associated with the structural and kinetic features of polymer melting. Typically the peak value is reported for polymer melting points. The instrumental error in temperature for a DSC is typically 0.5 to 1.0 C. Typical DSC trace for a Semi-Crystalline Polymer:

Glassy/Semi-Crystalline

Semi-Crystalline

Liquid

Scan Rate 10K/mi n Melting

Endothermic

Hysterisis

Cold Crystallization Glass Transition

Temperature (K)
The Figure above is a typical schematic for a heating run on a quenched sample of semi-crystalline polymer such as polyethylene, polyester (such as PETE) or isotactic polystyrene (typical atactic polystyrene does not display a crystalline endotherm). The left axis is (dH/dT), C p, or heat flux depending on the normalization of the heat flux. Also, the left axis is often plotted with the endotherm pointing down rather than up, flipping the curve. The curve can change dramatically with heating rate especially with respect to the hysteresis of the glass transition (residual enthalpy) and cold crystallization phenomena. The mechanical properties of the sample change from a brittle
3

solid such as polystyrene at room temperature at the left, to a typical semi-crystalline material such as polyethylene in a milk jug in the middle to a viscous liquid like molasses to the right. Determination of Tg : To determine Tg two lines are drawn parallel to the baseline above and below the inflection at T g. The midpoint of the departure from the left and the intersection with the baseline nearest Tg on the right is determined by the instrument. This midpoint is taken as the glass transition temperature. The residual enthalpy associated with Tg is sometimes also recorded from the high temperature baseline. This residual enthalpy will decrease with heating rate and has a strong dependence on the processing conditions of the sample and the length of time the sample is allowed to anneal near Tg. Tg is a second order transition since it displays a discontinuity in the derivative of the enthalpy (heat capacity). It is often termed a pseudo-second order transition because it displays a finite range of temperatures over which it occurs at finite heating rates and often displays a residual enthalpy. Determination of Tm: The melting point, Tm is determined, for broad melting polymers, by the temperature at the maximum in the (dH/dT) plot near this transition. The enthalpy of melting is determined by constructing a baseline above the melting and extending it to below any cold crystallization phenomena (exothermic peak below Tm and above Tg). Generally, the enthalpy of crystallization is taken as the area above this baseline. Sometimes the enthalpy of crystallization, which has occurred in the DSC measurement, is subtracted from this value to adjust the enthalpy as an estimate of the enthalpy associated with the original sample. The onset of melting is taken as the initial rise of the curve above baseline. The maximum melting temperature at this heating rate is taken as the final point which deviates from the baseline. Melting Endotherm: Gibbs-Thompson Relationship for Polymers (Hoffman-Lauritzen Theory): (Appendix 1) In addition to the broad melting endotherms and the presence of cold crystallization as a dominant feature in semi-crystalline polymers a number of unique complications exist in the interpretation of calometric data. A variant of the Gibbs-Thompson equation, known as the Hoffman-Lauritzen equation governs the relationship between structure and melting point for lamellar crystallites which are present in essentially all semi-crystalline polymers. The Hoffman equation notes that there is a direct, inverse relationship between the undercooling at which the polymer crystallites melt and the thickness of the lamellar crystallites through: L= 2 e Tm Hc T

(+L)

where T is the difference between the melting point of an infinite thickness and perfect crystallite and the observed melting point, e is the free energy associated with the lamellar fold surfaces and L is a term which accounts for the necessary deviation from equilibrium in the crystallization process. The latter term is usually neglected, i.e. L 0. The Hoffman approach gives rise to a possible description of the broad endotherm associated with melting of polymer crystallites, that is polydispersity in lamellar thickness. In many plastics, particularly in branched polyethylene, multiple melting endotherms are observed and the Hoffman relationship has been used to support the presence of discrete populations of crystallites associated, perhaps, with initial crystallization and formation of spherulitic structures, followed by decoration of spherulites by lower melting (i.e. thinner) lamellae.
4

A problem with the above analysis based on the Hoffman equation is that the degree of crystallinity (DOC) measured calorimetrically rarely agrees with that measured by XRD. XRD's DOC is usually taken as the best value since the myriad of complications which can be associated with a thermal measurement are mostly avoided. The value of the DOC obtained from thermal analysis, if all endotherms are included is usually higher than that of XRD which indicates that at least some of the endotherm measured calormetrically is associated with enthalpies of association present in the amorphous phase, such as tethered chains at the lamellar interface or interlamellar amorphous polymer. In some highly branched polyethylenes, it has been proposed that the bulk of the melting endotherm can be associated with such associated amorphous materials. Melting Point Depression for Polymers: (Appendix 1) The presence of miscible, non-crystallizing components are known to depress the melting point of all crystalline materials, i.e. adding salt to ice: 1 1 R ln( a) 0 = Tm Tm Hm where a is the activity of the crystallizable component. Often the mole fraction of crystallizable component is substituted for the activity as an ideal approximation. In polymers, the noncrystallizable component can be a low-molecular weight impurity, but is often a non-crystallizable comonomer, an atactic segement, end-groups or a branch site. Substitution of the crystallizable fraction for "a" in the above equation yields the Sanchez-Eby Equation for copolymers. For low fractions non-crystallizable component, such as when considering end-group effects the log term can be approximated using, ln(a) ln(Xcry) = ln(1-Xnoncry) X noncry. For end-groups X noncry = 2M0/Mn, where M0 is the molecular weight of the end-group and Mn is the number average chain molecular weight. Thus, T0 for a polymer reflects the melting point of an infinite molecular weight sample. For a thermodynamically miscible polymer or solvent at low concentrations the melting point depression can be expressed through a viral expansion where -ln(a) is replaced by a second order viral expansion involving the interaction parameter, : 1 1 R Vmer 2 0 = solvent solvent ( ) Tm Tm Hm Vsolvent where V represents molar volumes and is the volume fraction of solvent at low volume fractions. This equation is appropriate for melting point depression in the presence of a plasticizer for instance. The melting point of a semi-crystalline plastic sample is strongly effected by distributions in morphology (lamellar thickness), impurity concentration and distribution in the sample, and processing history (thermal history) as well as the residual strain and orientation in the samples. All of these effects lead to a characteristically broad melting endotherm for plastics. Glass Transition: The melting point is a first order transition, in the Ehrenfest sense, since it involves a discontinuity in the first derivative of the Gibbs free energy with respect to temperature (-entropy) or pressure (volume). That is, there is a discontinuity in volume, for instance, at the melting point. The glasstransition, in the most ideal case, is a second order transition since the first derivatives of the Gibbs free energy are not discontinuous but the second derivatives with respect to temperature (-heat
5

capacity/T), pressure (-volume * compressibility) and temperature/pressure (volume * thermal expansion) show a discontinuity. Polymers are unique in the dominance of the glass transition as the decisive factor in their mechanical properties. Polymers are the only material for which the equilibrium ground state is often glassy rather than crystalline. This is because topological and stereochemical constraints prevent the formation of crystals in many cases. The glass transition is often called a pseudo-second order transition because of the dominance of kinetics. Slower cooling rates in the DSC, for instance, lead to lower measured values of the glass transition. This time-temperature superposition is described by the Williams Landel Ferry (WLF) equation for example. This rate dependence hints at the basis of the glass transition in molecular motion. The glass transition is effected by orientation, rate, molecular weight, crosslink density and impurity content. Molecular Weight (Flory-Fox Approach): Through consideration of the free volume at the glass transition an expression can be obtained for the molecular weight dependence of the glass transition temperature which is based on the idea that chain ends lead to more free volume than mer units in the middle of chains: Tg = Tg, K Mn

where K is a constant (K = 2.1 x 105 for polystyrene, with Tg, = 106C) and Mn is the number average molecular weight. The term K/Mn is proportional to the number of end groups as in the melting point equation above. Polymer Blends (Fox Approach): The Fox equation describes the glass transition temperature of a miscible blend of two polymers, a copolymer or a plasticized polymer: 1 m m = 1 + 2 Tg Tg,1 Tg, 2 where mi is the mass fraction of polymer "i". The Fox equation leads to a lower value of Tg than would be given by a simple linear rule of mixtures and reflects the effective higher free volume or randomness due to the presence of two components in a mixture. Systems which obey the Fox equation are considered to display intimate and uniform mixing while those which deviate from it, especially those that display two glass transition temperatures are considered to be poorly mixed. Tacticity and Tm/Tg : Disubstituted vinyl polymers show dramatically different glass transition temperatures for different tactic forms (e.g. polymethylmethacrylate, PMMA: isotactic 43C, atactic 105C, syndiotactic 160C). Monosubstituted vinyl polymers show a single glass transition temperature for the different tactic forms (e.g. polystyrene, PS, 105C for all tactic forms). Melting points are generally dramatically different for different tactic forms since the different tacticities, i.e. isotactic versus syndiotactic, show different crystalline structures. Thermal Gravimetric Analysis (TGA): Many DTA instruments include the ability to measure mass as a function of temperature as well as the DTA output. In some cases such a thermal gravimetric analysis instrument is coupled with a mass spectrometer or an infrared absorption instrument for analysis of decomposition gasses.
6

Typically, a superimposed plot of Tsam-ref and weight can yield critical information concerning the changes which occur on processing a polymer. For instance, the thermal cycling of a processing operation can sometimes be mimicked in a DTA/TGA instrument to understand degradation and thermal transitions which effect the viscosity and other properties of a plastic. Often, a DTA/TGA analysis is used to define the processing limits for a polymer, at the lower temperature associated with the glass transition or melting point and at the upper temperature associated with degradation of the polymer. Polymers which absorb water, such as nylon, have been studied in depth using TGA instruments and an example of such a study is shown in figure 12.9 of Campbell and White. Since the TGA instrument is fairly simple and self-explanatory it will not be extensively discussed. Dynamic Mechanical Thermal Analysis (DMTA): When a polymer is subjected to a forced mechanical vibration at a fixed frequency, temperature and elongation a fraction of the energy is absorbed and a fraction is returned elastically. This is evident if a rubber ball is dropped on the floor. The ball bounces back to a height, L', lower than the height from which it was dropped, L. The loss associated with the ball bouncing can be quantified as L" = L - L', for instance. If one were to vary temperature or speed of impact the rebound fraction would change. This will be demonstrated using a rubber ball cooled with liquid nitrogen in class. In systems subjected to cyclic deformations or distortions, such as alternating current circuits, the loss term, L" is often associated with an imaginary component to a complex parameter, L*. So we would write, L* = L' + i L" where i is (-1). In complex space, a 2-D plot with the x-axis real and the y-axis complex, the complex parameter L* is represented as a point and the angle, , from the real or x-axis to a line from the origin to the point L* is given by: tan = L"/L'. tan is a "normalized" measure or the mechanical loss associated with the bouncing ball at a fixed temperature and frequency (speed of impact). Any mechanical or rheological measurement can be conducted in a dynamic mode, i.e. with a sinusoidally oscillating force. For example, a tensile stress experiment involving an oscillating force, F() = F0 sin(2t) where F0 is the amplitude of the force, is the frequency and t is the time in the inverse units of , gives rise to a dynamic stress, () = F()/A = 0 sin(2t), where A is the sample's cross sectional area. The resulting strain, = L()/L0, for a viscoelastic material will display an in-phase component and an out of phase component, = ' sin(2t) + " cos(2t). More typically, the strain is the independent parameter, = 0 sin(2t) and the stress is a complex response, since the material response is often a function of the maximum strain, 0. A complex Young's modulus can then be constructed, E* = E' + i E", with an associated reduced loss, E"/E' = tan. Similarly, a complex viscosity can be constructed for a shear experiment. Figure 12.10 in Campbell and White shows several arrangements for dynamic mechanical experiments. Figure 12.11 in Campbell and White shows a typical measurement involving a glass and a crystalline transition for a semi-crystalline polymer. Below the glass transition the polymer is glassy and the elastic response is similar to a glass marble, i.e. E" is small and E' is large so tan = E"/E' is small. Above the glass transition the material is elastic and E" is small and E' is smaller. tan is slightly higher but lower than at Tg. At Tg, molecular motion leads to significant absorption
7

at the frequency of the dynamic strain leading to a large E". E' is continuous through the transition so a peak in tan at the glass transition results. In class we showed that a rubber ball behaves like a piece of lead at the glass transition, i.e. a highly lossy material. The real modulus, E', of the semicrystalline material in Figure 12.11 decays with temperature above the glass transition until the crystalline melting point where the material becomes a liquid. tan shows a monotonic increase until it peaks at the crystalline melting point due to enhanced molecular motion when crystals begin to melt. In many polymers a number of absorption peaks are observed below the glass transition which can be associated with different types of molecular motion. The glass transition in this scheme is termed the primary transition temperature or -transition, T, and the secondary transitions are labeled in sequence of decreasing temperature the , , and transitions with associated temperatures T for instance. One of the uses for DMTA is in understanding the mechanical response of elastomers such as those used in tires. In a tire high frequency (low-temperature) loss is often considered good since it enhances the grip of a tire at high speeds (high frequency). Low frequency (high-temperature) loss is sometimes considered bad since it increases wear and reduces gas mileage. Elastomers can be tuned to broaden the mechanical absorption peak in certain temperature ranges by controlling chemical composition, chemical structure, and the structure of fillers such as carbon black which can make up more than half of the weight of a tire. The mechanical absorption spectrum is also critical to a wide range of plastics since the mechanical behavior of polymers varies with both temperature as well as speed of deformation through the time-temperature superposition principle. Time-Temperature Superposition: In class a sample of silly-putty was used to demonstrate time-temperature superposition. If silly putty is left for a long period of time (minutes) it flows (liquid at long times, low frequency or high-temperatures). If a ball is made it will bounce (seconds, elastic at intermediate times, frequencies and temperatures). If silly putty is rapidly ripped apart it fails like a glass (glassy at very short times, high-frequency or low temperatures). From this observation we can consider that raising the temperature is similar to dropping the frequency or allowing more time for deformation. Williams-Landel-Ferry (WLF) considered the equivalency of time and temperature in the context of free volume theory for an activated flow process in viscoelastic materials. The WLF equation yields an equivalent frequency for a given temperature relative to a ground state temperature and experimental frequency: ln aT = C2 + (T1 T0 ) C1 (T1 T0 )

0 = aT

where C1 and C2 are constants for a given polymer and T0 is a reference temperature for a given polymer close to Tg. The constants in the WLF equation can be theoretically predicted from freevolume theory (C1 = 17.44 and C2 = 51.6, using T0 = Tg) or can be experimentally determined. Using the WLF equation a master curve can be constructed in temperature that corresponds to a standard frequency or a master curve in frequency can be constructed that corresponds to a standard temperature.

Ideal Models in the DMTA Measurement: For an applied dynamic strain, = 0 sin(2t), an ideal Hookean elastic will respond with a complex stress completely in phase with the applied strain: * = 0 sin(2t) = E 0 sin(2t). For a Hookean elastic E" = 0 For a Newtonian Fluid, an ideal viscous fluid, the stress is given by: = (d/dt). d/dt for the dynamic strain, = 0 sin(2t), is given by: d/dt = 2 0 cos(2t), so, for an ideal lossy material, * = 2 0 cos(2t), and E' = 0. DMTA experiments can be performed on non-Newtonian fluids such as a polymer melt to determine the elastic component of a visco-elastic fluid or on a solid plastic or rubber to determine the viscous component of a visco-elastic solid. Work of Dynamic Deformation: The work performed in a DMTA measurement per unit sample volume per cycle, W, is given by: W = * d * = 0 0 sin( ) = ( 0 ) E"
2 0 2

The power consumed per unit volume is given by, P = ( 0 ) E"


2

DMTA Measurement of Complex Shear Modulus: One of the most common applications of DMTA measurements is the determination of the shear loss and storage modulus, G" and G' for a polymer melt. This measurement is critical to the understanding of polymer processing since a polymer melt is subjected to a variety of shear rates in a typical process. For example, in an extruder at early stages low shear melts polymer pellets and pressurizes the polymer fluid in the feed and melting zones of the extruder barrel. In the extruder die this pressurized melt is subjected to extremely high rates of shear as it is formed into the final extruded part. The energy consumed in the process is related to the complex shear modulus and the variation of the loss shear modulus with rate. For a dynamic shear strain,
9

= 0ei2t = 0 sin(2t) where the shear strain is defined as the change in length of a fluid with respect to a normal direction. The complex shear stress, force per normal area dragging the fluid is given by Newton's viscosity law by,

* = *

d * = i 2 * * dt

The complex viscosity is written as,

* = ' i"
(compare with complex modulus which is written E* = E' + iE"). The difference between this equation and that for the complex Young's Modulus occurs because the complex viscosity involves the measurement of the compliance of the fluid to an applied strain rate rather than the strain rate of a material due to an applied stress. The compliance, D, is the inverse of the modulus, E. For complex numbers the inverse is obtained by the ratio of the complex conjugate to the magnitude of the complex number: D* = 1 conj ( E*) E' E" = = D' iD" 2 = 2 i 2 2 2 E * ( E' ) + ( E") ( E' ) + ( E") ( E' ) + ( E")2

The shear loss modulus, G", and shear storage modulus, G', are related to the dynamic viscosity * = ' - i" by, G' = 2" G"= 2'

The Cox-Merz Rule gives the zero shear rate viscosity, 0, as the magnitude of the complex viscosity,

0 = * = (' ) + (")
2

2 Qualitative Examples of the Use of DMTA in the Plastics Industry: 1) In the processing of high density polyethylene a dramatic increase in the processing cost is associated with small amounts of long chain branching. This can be at levels below 5 branches per 1000 carbons in the main chain so can not be characterized using spectroscopic techniques. Rheological DMTA has been used to estimate the amount of long chain branching since the highfrequency shear loss modulus, G", displays a characteristic increase relative to the intermediate frequency loss modulus for branched materials. Presumably this effect is associated with hindrance of molecular motion at high frequencies due to the presence of long chain branches. There is good correlation between this measurement and the processing costs as well as some quantitative correlation between branch content and this mechanical response. DMTA is the only analytic technique which has been demonstrated to be sensitive to such weak chain branching and it has proven critical to control of synthesis conditions aimed at reduction of long chain branching in HDPE. 2) In the elastomer industry DMTA is a standard instrument for determination of the expected performance of rubber products. The most common example is a automotive tire which is subject
10

to a wide range of dynamic strains in use. The frequency dependence of loss in elastomer compounds can be directly related to the performance of an elastomeric material in an automotive tire. The critical design criterion, in terms of materials, for a tire are "grip", "wear" and "fuel economy". The "grip" of a tire is usually associated with the high-frequency (low temperature) loss. The low-frequency (high temperature) loss is associated with the wear and loss of fuel economy. The addition of carbon black to an elastomer enhances the high frequency (low temperature) loss, leading to higher grip. The weight fraction of carbon in a tire is typically between 40 and 60%! In carbon reinforced elastomers, this higher grip is always associated with an increase in the low frequency response leading to an increase in energy absorption or reduced fuel economy for a tire as well as an associated increase in wear. The increase in loss with addition of carbon has been related to the association of elastomer molecules with filler, so called "boundrubber content". Loss is also associated with breakup and re-formation of the loosely associated aggregates of carbon. In the design of elastomers for tire applications the goal is often to broaden the absorption peaks in the DMTA spectrum and to shift absorption's to higher frequencies through modification of the chemical structure of the elastomer and tuning of the filler structure. Appendix 1: Gibbs-Thompson Equation and Melting Point Depression Equations At the equilibrium melting point a crystal and its melt have equal Gibbs free energy, Gcrystal = Gmelt. For example, in a glass of water with ice cubes at close to 0C the ice will fuse into a larger cluster due to epitaxial recrystallization (crystal surface nucleated crystallization). This occurs because there is little or no difference in free energy between water molecules in the liquid and crystalline states. An equilibrium melting point is defined as: G = 0 = H T S or T = H S

where is the difference between the crystal and the melt and T reflects ideal conditions. There are many ways to deviate from this ideal situation. For example, the ideal crystal above is of infinite size. For real crystals, where surface area becomes important at small crystallite sizes the crystal will display an equilibrium with its melt at a different temperature due to the inclusion of a surface energy term: VG = 0 = VH TmVS A s or Tm = VH A s AT = T s VS VH

A melting point depression is predicted for a finite sized crystal depending on the surface to volume ratio and the surface energy to bulk enthalpy ratio. For low surface energy crystals or for large crystals the effect is smaller. For polymers, crystallites are highly asymmetric with roughly the aspect ratio of a sheet of paper. This means that the A to V ratio is very large. Additionally, the planar surface of polymer crystals contain chain folds which have a high tortional energy associated with them. This means that the surface energy to enthalpy ratio is also large. Large under coolings are expected in polymer crystals and a direct relationship between the lamellar thickness and the melting point is predicted by the Gibbs Thompson Equation given above. For a system such as a collection of red and blue marbles which are shaken up the entropy of the system can be calculated from combinatorial (or counting) statistics. The difference in entropy between a segregated state (all blue at the top and all red at the bottom) and a totally mixed state (blue and red randomly dispersed) is given by the total number of ways to arrange the marbles in the dispersed state, :
11

S = R ln() If an impurity is added to a crystal and it has a totally uniform dispersion then a change in entropy between the pure crystal and the crystal with the dispersant can be calculated form a combinatorial approach which parallels the red and blue marble law above assuming that the number of states increases with the molar concentration of the impurity, x: S = R ln( x ) Using an approach similar to the Gibbs Thompson discussion above, a melting point depression can be predicted: G = 0 = H Tm S RTm ln( x ) rearranging: 1 1 R ln( x ) = + Tm Tpure H or Tm = H S + R ln( x )

12

Chapter 4. Electromagnetic Radiation in Analysis (Chapter 3 Campbell & White). Electromagnetic radiation is a disturbance in electro-magnetic space which follows Maxwell's differential equations for conversion of energy from an electrical field to a magnetic field. The disturbance is sinusoidal in nature for propagating EM radiation. Electromagnetic waves display a wave nature in that the oscillating electric and magnetic fields propagate with a wavelength and frequency. Additionally, a special property of EM radiation is that it displays a fixed speed in vacuum, c , and fixed velocities in other uniform media so conversion from frequency to wavelength is direct: =c / We can specify a type of EM radiation by specifying it's wavelength, frequency, or wavenumber, k: k = 1/ EM radiation also displays a particle characteristic through the concept of the photon which is a particle of no mass. This allows a means to describe features of EM radiation which are usually associated with particles such as momentum. The energy per photon of EM radiation is related to the frequency, , of the Maxwellian sinusoidal oscillation through Planck's constant, h: Energy = h That is, higher frequency is associated with higher energy. Then the Energy is given by hc/ and different wavelength radiations contain different amounts of energy per photon. A photon is a quantum of EM radiation that displays momentum. The momentum is expressed as: p = h/ =h/c The brilliance, brightness, flux or intensity of a particular EM radiation is related to how many photons are delivered in a unit area per unit time. The energy of each photon is related directly to the wavelength. It is important to be able to distinguish in your mind between intensity, energy and power delivered by EM radiation. The energy of EM radiation determines the effect of the radiation on materials and is the basis for understanding analytic techniques which use EM radiation. The intensity determines the amount or number of these effects which occur. For example, x-rays are of short wavelength and have very high energy per photon. Because of this x-rays are a type of ionizing radiation which can cause cancer in humans through formation of free radicals in your DNA for instance. Radio waves are of very large wavelength and low energy. Radio waves pass through our bodies constantly with no effect. High intensity radio waves can cause changes in the spin polarization of nuclei when the nuclei are under a very strong magnetic field and this is the basis of nuclear magnetic resonance absorption which will be discussed in class.

Analytic Techniques Using EM Radiation: (Low Energy to High Energy) NAME Nuclear Magnetic Resonance (NMR) Electron Spin Resonance (ESR) Microwave Absorption Infrared Absorption IR Raman Spectroscopy Elastic Light Scattering Inelastic Light Scattering UV absorption Elastic X-ray Scattering XRD, SAXS Inelastic X-ray ESCA Ionization (SIMS) 10 -6 -10 - 4 10 -9 -10 - 6 Gamma-Rays Cosmic Rays Ionizing Ionizing 0.4-0.7 Far IR IR Near IR Wavelength Common Name m 10 1 0 Radio 107 2x105 15.4-830 2.5-15.4 0.7-2.5 TV Radar IR IR Visible Light UV-A, UV-B X-Rays Changes in Polarization of Molecules Move electrons in Orbitals Ionizing Effect Nuclear Spin Electron Spin Rotate Polar Molecules Bond Vibrations

0.01-0.4 10 -4 -10 - 2

Features of Electromagnetic Waves: 1) EM waves are described using a sin function in time and space. That is, at a fixed time the wave varies in Amplitude, A , with spatial position and for a given position in space the wave varies in amplitude with time. This is true of both the electric field vector, E , and the magnetic field vector, B. E = A sin[2x/ -2t] B = B0 cos[2x/ -2t] Out of phase with E by 90 E and A have direction. is the wavelength, is the frequency. 2) The frequency and wavelength are related by the velocity of the EM wave which in vacuum is c, "the speed of light". = |c|/ c = 3 x 108 m/s 3) The amplitude is related to the "intensity" of the EM radiation: Intensity = |A|2
2

Intensity doesn't have direction, amplitude does. 4) The energy of EM radiation is directly related to the energy associated w i t h phenomena by which it is made and is not associated with the intensity. Planck's law states: Energy/photon = h = hc/ h = 6.63x10-34 J s The idea that EM radiation has an energy is tied to the particle view of EM radiation, i.e. there is a particle called a photon which has no mass which carries the EM energy and has momentum, p. p = h/c = h/ (c/c) 5) Light, X-rays and other EM radiations are generally composed of a number of photons so one can consider the relationship between different waves in a beam. i.) Coherent: If all waves in a beam are in-phase , that is have the same phase angle (peaks of waves coincide in space) they are called coherent . Waves which are not coherent can interfere with each other leading to a reduction of the intensity. For example, a laser beam is coherent while a flash light beam is incoherent. This is one reason why a laser beam can propagate over great distances while a flash light beam quickly dissipates. When considering interference of two waves one adds or subtracts amplitudes of the electric field vector E . The intensity which is measured is the square of the resulting amplitude, #3 above. ii.) Collimated: Beams with waves which are all progressing in the same direction are termed a well collimated beam. Collimation refers to the divergence of the waves in a beam. A light bulb produces uncollimated light which spreads in all directions. The sun's rays, when they reach earth are well collimated since the angular divergence is low. A laser beam or a synchrotron x-ray beam are well collimated due to the mechanism by which the EM radiation is produced. iii.) Monochromatic: If all waves have the same frequency (or wavelength by #2 above) they are called monochromatic (one color). A source like a light bulb, the sun, or an x-ray tube generates polychromatic (white light) radiation (many wavelengths) and a source like a laser or an x-ray synchrotron yields monochromatic radiation. The polychromacity of EM radiation is tied to the mechanism of formation. If the formation event is specific (quantum) in terms of the energy transfer associated with the formation event, monochromatic radiation results. If the formation event is statistical (distributed in energy) polychromatic radiation results. iv.) Polarization: The electric field vector, E , for and EM wave has a direction in a plane normal to the propagation direction. If the direction is fixed relative to the propagation direction and this direction is the same for all waves in a beam, the EM radiation is said to be linearly polarized. Polarization can be produced by a number of means: Reflection off a surface leads to linear polarization in the plane of the surface, highly birefringent materials can lead to polarization of a beam by absorption of components not with a certain polarization, some processes for formation of EM radiation produce polarized radiation laser and x-ray synchrotrons, in some cases a grating can be used to polarize radiation (Soller slits for x-rays), diffraction leads to polarized radiation depending on the geometry of diffraction.

In addition to linear polarization, waves can be elliptically and circularly polarized. In circularly polarized beams the vector E rotates in direction along the propagation direction. Elliptically polarized radiation is a mixture of circularly and linearly polarized radiations, i.e. there is some rotation of the vector E but it is not symmetric. In this class we will only discuss unpolarized and linearly polarized radiation. 7) EM radiation can always be assumed to travel in a straight line. 8) EM radiation interacts with matter in different ways depending on the energy associated with a photon. That is, energy decides what happens while intensity decides how much happens. Radio waves are low energy/high wavelength (see table above) and can pass through most materials with no effect. IR vibrates bonds and can generate heat. Light changes the polarization of molecules which is a minor effect. UV can dissociate weak bonds and cause degradation. X-rays are a type of ionizing radiation that can ionize atoms and molecules. Typically, x-rays have wavelengths on the ngstrom scale. Generally, the lower the wavelength of radiation the higher the danger due to the higher energy associated with short wavelengths see #4 above. It is also more difficult to produce and use high-energy photons since they must result from an associated high energy event and they are absorbed by most materials through the interactions mentioned above. 9) For visible light the index of refraction, n, is used to describe the speed of the radiation in a medium. Velocity = v = c/n. Since the velocity is always smaller than the velocity in vacuum, n is always greater than 1. In air, n is close to 1, in silica glass n = 1.52, some materials have a very high index of refraction, Titania, TiO2, is close to 5. Beer-Lambert Law: As a first approximation all EM radiation can be considered to follow a linear absorption behavior described by an exponential decay of intensity with thickness of a sample and with concentration of the absorbing species. This is critical to quantitative analysis using EM absorption which is the basis for IR, NMR, and UV spectroscopies. The Beer Lambert Law states that the value of the absorption is proportional to the amount of the absorbing species. In a typical absorption experiment, a beam of EM radiation passes through a sample of thickness "t". For a small thickness, dx, the incident intensity at wavelength and position x, I(x), is reduced by dI due to absorption by the sample. The change in intensity is proportional to the thickness, dx, the incident intensity I(x) and the linear absorption coefficient, A: dI = AI ( x )dx = cI ( x )dx For a thicker sample this equation can be rearranged and integrated: dI = I I ( x = 0)
I( x =t) x =t

x=0

cdx
or I = T = e ct I0
4

I ln = ct I0

"T" is the transmission ratio for the sample at wavelength l, and A = c is the Absorption. The Beer-Lambert Law given above gives t A = ln(1/T). Absorption spectra are either plotted as transmission or as absorption on the y-axis. In a transmission plot the absorption peaks point down. In an absorption plot the absorption peaks point up and are the ln of 1/T as above. Scattered Intensity: For a scattering experiment such as XRD, SAXS, or Raman scattering the Intensity is plotted versus either and energy or angular term. The integrated intensity for a scattering peak is directly proportional to the amount of the scattering material. A Raman scattering experiment results in data which looks very similar to an IR absorption experiment with both plotted versus wavenumber = k = 2/. IR will be a plot of Transmittance (peaks down) or Absorption (peaks up) and a Raman pattern will be a plot of Intensity (peaks up). Instrumentation for Spectroscopy: Dispersive Spectrometer: Figure 4.1 of Campbell and White, pp. 44, shows a typical dispersive optics spectrometer (UV in this case). Even in this simple instrument the optical paths are quite complicated. The first requirement of a spectrometer is resolution of the spectrum of an source through a tunable device which can isolate a single wavelength radiation. In a dispersive instrument this can be done with a prism or with an optical grating. The prism disperses the incident radiation due to differences in the index of refraction of the glass prism with wavelength, n( ). Snell's Law can be used to determine the angle of refraction of different wavelengths: nAir sin(Air) = nGlass sin(Glass) where Air is the angle of the incident beam with respect to a normal to the surface of the glass, and Glass is the angle of the beam in the prism with respect to the glass surface normal. The double refraction of the prism serves to disperse the incident beam into angularly diverging beams of different wavelength. The prism can be rotated to select certain wavelengths using a slit. An optical grating can also be used to disperse wavelengths using Bragg's Law: sin()=/2d where 2 is the angle of divergence from the incident beam. The main problem with dispersive instruments is that almost all of the incident radiation i s lost in the slit arrangement for selecting a diverging beam of the desired wavelength. The measurement of a spectrum involves rotation of the prism or grating. Most spectrometers involve double beam optics (shown in Figure 4.1) because, 1) the source has a spectrum of its own, 2) imperfections in the prism or grating might selectively effect certain wavelengths, 3) the detector usually has a spectral sensitivity which varies with wavelength and 4) the instrumental sensitivity might vary in time due to atmospheric interference and fluctuations in the electronics. Double beam optics are achieved by a partially silvered mirror which effectively splits the incident beam into two beams, a Reference beam and the Sample beam. A reference cell is used which
5

duplicates the sample except that no sample is present. A rotating mirror is used to alternatively sample the reference and sample beams and the transmission is measured as the ratio of these two beams using a photomultiplier (PM) tube which results in a number of counts or events. The raw counting error for such a spectrometer is related to the square root of the number of counts measured by the PM tube. Usually, this counting error is much smaller than the error introduced by the instrumental setup such as the spectral width attainable with the slit/prism optics. Fourier Transform Instruments: The loss of information due to the dispersion/slit optics in a dispersive spectrometer was long known to be a major hindrance to the determination of EM absorption spectra. With the development of the Michelson Interferometer a route to alleviate this problem became clear. Figure 5.2 of Campbell and White, pp. 57, shows a schematic of a Fourier Transform IR absorption spectrometer. The Fourier transform instrument results in an interference pattern between two EM beams which is the Fourier equivalent of an absorption spectrum. Fourier transform mathematics are critical to IR and Raman spectrometers, NMR spectroscopy, and is the basis for scattering techniques such as XRD, SAXS and SALS we w i l l spend some time discussing FTIR instruments as a simple introduction to Fourier techniques in Polymer Analysis.

Detector

Sample

Movable Mirror I

x x

Fixed Mirror
Consider the instrument shown schematically above. Polychromatic radiation (see definition above) is incident on a sample and the intensity versus wavelength spectrum is modified by absorption of radiation at certain characteristic wavelengths by the sample (see below). A polychromatic spectrum is composed of a series of EM waves of different wavelengths, with different amplitudes. This continuous distribution of wavelengths can be decomposed into the plot
6

shown below, i.e. Intensity = |A2| versus wavelength. That is, the intensity versus wavelength plot is a population distribution plot for these sine waves. A basic premise of Fourier Theory is that the components of this distribution can be treated independent o f the other components. This approach allows us to consider a single sine wave from the polydisperse distribution, and recombine the separate sine waves into the polychromatic spectrum after considering interference and absorption for each wavelength independently.

Intensity

Sample

Wavelength,

Intensity

Wavelength,

Consider a single monochromatic sine wave which enters the double mirror at the center of the FTIR schematic above. Identical sine waves propagate to the movable mirror and to the fixed mirror. If the distance traveled to the fixed mirror is L then the path length of the Fixed Mirror beam is 2L. The movable mirror is positioned at location L+x from the central double mirror. The path length for the movable mirror on return to the double mirror is 2L + 2x, and the path length difference for the two beams at the double mirror is 2x. The two beams, one from the Fixed Mirror and the other from the Movable Mirror, are recombined as they progress directly upward in the schematic. The combination of these two beams allows for interference to occur between the two beams, i.e. the phase difference 2 (2x/) will lead to some conditions of destructive interference between the two beams (Amplitude of one will be positive and the Amplitude of the other will be negative). Summing the amplitude will lead to a smaller total amplitude for most cases, and this reduction in amplitude will be a cosine wave in 4x/, with a maximum intensity = |A2| at x = 0.

1.0

Amplitude

0.5 0.0 -0.5 -1.0 -40 -20 0 20 40

x
This function is the Fourier transform of a single wavelength, monochromatic, distribution. That is, if we decompose this function into a series of sine waves we find that the function is represented as a single value in the amplitude versus wavelength plot.
7

Fourier transformation results in inversion of the units of the independent a x i s . This means that the transformation of x will result in a distribution in w a v e number, i.e. 1/x =k. For a monochromatic wave, the interferometer results in a cosine wave centered on x=0. This would result if a laser were incident on the double mirrors, for instance. For a polychromatic wave a series of amplitude versus x plots would result with different wavelengths, across the spectrum of Intensity versus Wavelength shown above. The amplitudes of these polychromatic waves sum after passing through the optics of the interferometer. The resulting sum is a damped cosine wave in x with a central peak at x=0. The decay of this cosine wave is not monotonic, i.e. If the amplitudes at the peaks are plotted against x there is no simple function which describes the decay. The decay contains all of the information concerning the spectra incident on the interferometer. The information, however, can not be directly interpreted from the decay pattern since it is in " inverse space ". Inverse space refers to the x-dimension here. This is inverse in units to the wavenumber, k. Real space in this experiment is wavenumber. These two spaces are related by a Fourier transform.

Amplitude

Fourier Transform 0 x

Intensity

Wavenumber, k=1/

The biggest advantage of a Fourier transform IR instrument over the dispersive instrument discussed above is that all of the incident wavelengths are used in determining the spectrum, i.e. none of the incident radiation is disposed of. In the dispersive instrument almost all of the incident radiation is removed by the slit, see above. The narrowest resolvable feature in wavenumber space is much smaller in a FT instrument. There are a number of other more complicated advantages to the Fourier Transform instrument which will not be discussed here. Dispersive instruments are cheaper since they do not require a computer, a sub-micron translation stage, i.e. x has to be on the scale of the wavelength of light, and a number of other optical components needed in the FTIR instrument. There is also some advantage to a dispersive instrument when the interest is on a single or very narrow range of wavelengths. Pages 57-58 of Campbell and White discuss the FTIR instrument. The governing equation for the Fourier transform in FTIR's is given there as, G( ) =

I ( x ) cos(2x )dx

(note error in the equation in Campbell). G is the intensity in frequency or wavenumber space and I is the intensity measured as a function of x. This integral states that the Intensity as a function of is a sum of cosine waves of frequency in x space. Fourier transforms are invertable so we can also say that this equation implies that the sum of intensities in frequency space at a position x yield the intensity as a function of x. The basic idea is that intensity can be decomposed into independent sine or cosine components of different amplitudes in either x or frequency space.

Fourier Transforms in Other Analytic Techniques: Several other analytic techniques rely on either Fourier Transform instruments or Fourier Theory to measure analytic features. The most common are NMR, XRD and SAXS. In NMR an oscillating decay of radiowave emission in time results from a radiowave pulse in time. This is similar to the decay in tone of a guitar string after being plucked. Fourier transform of this time signal results directly in a frequency space spectrum. This will be discussed in detail when we discuss NMR. In Scattering techniques such as XRD and SAXS, measurements are made in angular space which can be converted to wavevector or "q" by q=4/ sin() where is half the scattering angle. The units of q are inverse size and a Fourier transform of the intensity in q-space results in the radial distribution function in real space (size space). We will discuss this when we discuss small angle scattering. This topic is also covered in Cullity in his Appendix on inverse space. (note Cullity uses "s" rather than "q". These two quantities are related by a constant factor of 2.)

PDF File: (Click to Down Load): Chapter5.pdf

Polymer Analysis
= Back to TOC = To Syllabus

Chapter 5. IR Spectroscopy and Raman Scattering


(Chapter 5 Campbell & White). Bristol University IR Spectroscopy Whitworth College IR/NMR problems Scimedia on spectroscopy CSU on IR http://chipo.chem.uic.edu/web1/ocol/spec/IR.htm (Application/Chemistry) http://www-wilson.ucsd.edu/education/spectroscopy/spectroscopy.html (Physics) http://www.columbia.edu/cu/chemistry/edison/IRTutor.html (Spectra) Good Movies http://avogadro.chem.iastate.edu/chem572/ (Chemistry) Introduction: The energy associated with electro-magnetic radiation in the infrared range (just above visible in wavelength) is sufficient to excite vibrations of chemical bonds. IR spectroscopy and Raman scattering both involve IR wavelength radiation and both characterize vibrations of chemical bonds. For this reason they are usually considered as a group although the instrumental details for the two techniques are significantly different. The vibration of any structure is analyzed in terms of the degrees of freedom which the structure possesses. For example, a sphere has 3 degrees of transitional freedom and 0 degrees of rotational freedom since rotation does not result in a perceptibly different state. A single sphere or a single atom does not have vibrational states. When two spheres are bonded the group has 3*2 degrees of translational freedom. The grouping of 2 spheres, as a unit, possesses 3 degrees of translational freedom and 2 degrees of rotational freedom, since rotation about the axis of the two spheres does not result in a perceptible change. When considering vibrational states we fix the reference frame on the grouping of objects, so the degrees of freedom for the grouping are subtracted from the total number of translational degrees of freedom for the individual spheres. For two spheres there are 3*2 - (3+2) = 1 degree of vibrational freedom. This means that the IR/Raman spectra for a diatomic molecule such as CO will have one absorption band. This vibration would involve stretching and compressing of the CO bond. For groupings of spheres with more than 2 members the number of vibrational states is 3n-6 so a molecule such as water has 3 vibrational states which results in three absorption bands in IR and Raman. These are symmetric stretching of the H-0 bonds, asymmetric stretching of the H-O bonds and a scissors bending of the HOH structure. The number of stretching vibrations is n-1 and the number of bending vibrations is 2n-5. Carbon dioxide, O=C=O, is a linear molecule so the number of degrees of freedom are 3n-5 rather than 3n-6, i.e. one of the molecular rotational degrees of freedom, rotation about the molecular axis, does not result in a perceptible change. The number of IR and Raman absorption bands is calculated from the number of degrees of translational freedom for the collection of atoms in a molecule, 3n, minus the number of degrees of translational and rotational freedom for the molecule as a whole, usually 6, but 5 for a linear molecule. These normal modes of vibration are useful for consideration of relatively small molecules, i.e. Benzene (C6H6) has 30 absorption bands in IR and Raman, each of which can be described in detail. Consider a polymer molecule such as a 100,000 gm/mole sample of polystyrene. This molecule contains about 1,000 mer units or 16,000 atoms! The number of vibrational states for this molecule are 3*16,000 - 6 or about 50,000 different vibrations. It is impossible to identify all of the vibrational states for such a molecule. In polymer analysis we can greatly simplify the characteristic spectra for a chain by considering the repeating chemical groups which occur in the chain as independently contributing to the IR and Raman spectra. This approach is called the group contribution approach, and for polystyrene, would involve consideration of the major bands due to aromatic ring bands, C-H vibrations, C-C and C=C vibrations. In the group contribution method many of the weaker bands are simply ignored and we concentrate on the few high absorption bands which serve as a finger print for a particular polymer.

IR Active Bands: The possible vibrations of a molecule are sensitive to IR absorption if the vibration results in a change in the dipole moment, u, of the molecule. The dipole moment is the product of the charge times distance and is similar to the moment of inertia in mechanics except that charge is the weighting factor rather than mass. When an EM wave in the IR wavelengths irradiates a molecule the electric field acts on the charge distribution in the molecule, i.e. the more polar the molecule the larger the effect. The oscillation of the EM electric field, if of the quantized frequency for absorption by a particular bond, will set the bond in motion, vibrating at the specific frequency needed for that vibrational excitation. The IR absorption experiment involves the oscillating electric field changing the charge distribution so that a dipole is enhanced or diminished. Strong IR absorption bands occur for polar groups such as OH, Cl, and the C=O bond. In determining if a vibration is IR active consider if there is a change in the sum of the charge*distance vectors, i.e. for a symmetric stretch of O=C=O (linear molecule) the movement of the left O is offset symmetrically by the movement of the right O so there is not net change in the charge * distance vector, not change in the dipole moment so this is not IR active. For a non-linear molecule like HOH (shaped like a V) there is a change in the dipole moment for a symmetric stretch so the vibration is IR active. Raman Active Bands and The Raman Scattering Experiment: The Raman scattering experiment involves shifts in the wavelength of an incident monochromatic beam. Raman scattering uses a laser as a light source (IR uses a mercury lamp or other broad spectrum source). The laser is usually in the optical wavelengths. The incident light causes motion of electrons in bonds. These moving electrons reemit light of the same wavelength for elastic scattering. Light scattering is sensitive to the mobility of electrons in bonds. The mobility of electrons in a bond is called the Polarizability of the bond, i.e. a measure of how easy it is to move electrons and polarize a bond. For bonds with a strong dipole moment (which are IR active) the mobility or polarizability is usually low. For bonds which have a weak dipole moment (which are IR inactive) the polarizability is usually high and the vibrational states of the bond are Raman active. IR and Raman activity are complimentary and the two techniques are used to fully characterize the vibrational states of molecules. Raman scattering is based on a scattering event as described above. Figure 5.4 of Campbell and White shows a schematic of a Raman spectrometer. A laser (usually an argon laser) is incident on a sample. Scattered light is collected usually at 90deg. to the incident beam. The spectrum of the scattered light is measured using either a dispersive spectrometer or a Fourier transform spectrometer. Small shifts in the wavelength from the incident wavelength due to inelastic scattering are measured in the Raman spectrometer. In the Raman measurement, an incident EM wave induces polarization of a bond through the EM wave's electric field. The energy of this excitation of the atom is h[nu]0, where [nu]0 is a larger frequency (and energy) than the IR range, [nu]vibration. There is a possibility that some of the energy of this excited state can be transferred to the atom in terms of a vibration of the bond, h[nu]vibration. Since the source of this energy is a scattering event, the absorption will be stronger for more polarizable bonds, i.e. bonds with more freedom of movement for the electrons. The loss of energy for the scattered EM wave due to transfer to a vibrational state for the bond is called a Stokes event and the resulting scattered wave is of higher wavelength and lower energy, EStokes = h[nu]0 - h[nu]vibration. The elastically scattered beam of energy ERayleigh = h[nu]0 is 10,000 times more intense than the Stokes line. It is also possible for an incident EM wave to interact with a bond which is already vibrationally excited. In this case an Anti-Stokes line of higher energy results, EAnti-Stokes = h[nu]0 + h[nu]vibration. The anti-Stokes line is much weaker than even the Stokes line. If a number of absorptions occur for a material, then the spectral distribution of scattered light can be measured and the shifts from the Rayleigh (elastic scattering) line converted to wavenumber. A spectrum very similar to an IR absorption spectrum results. In determining if a vibration is Raman active consider if there is a change in the volume of the electron cloud, i.e. for a symmetric stretch of O=C=O (linear molecule) the movement of the left O is in the opposite direction of the movement of the right O so there is a net change in the volume of the electron cloud within the molecule, this vibration is Raman active. For the asymmetric stretch the movements of the two O's are in the same direction so the volume increase on the left is offset by a volume decrease on the right and the asymmetric stretch is not Raman active. Since the source of the Raman spectrum is a scattering event, the Scattered Intensity is directly proportional to the concentration of species giving rise to the Raman lines. This is different than in an IR absorption experiment which follows Beer's Law discussed in the previous section. The following two spectra compare the IR and Raman absorption from nylon-6,6: {-(CO)-(CH2)6-(NH)-(CO)-(CH2)4-(CO)-}n. NH stretch is the highest wave number absorption. This is a polar bond so strongly absorbs in IR and weakly in Raman. CH stretch is a doublet below 3000cm-1 (asymmetric and symmetric) which is less polar than NH and has a strong absorption in Raman but a weaker absorption in IR. Carbonyl stretch is a signature band (about 1750cm-1) in IR which is weak in Raman (highly polar and volumetrically inflexible bond).

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993. Intensity and wavenumber of absorptions in IR and Raman The wavenumber (energy or frequency) of an IR/Raman Absorption depends on the mass of the atoms connected to a chemical bond, the strength of the chemical bond, and the geometry of the molecule. There are two basic types of vibrations, Stretches and Bends. Bends require less energy so occur at lower frequencies for the same or similar bonds. There are generally two types of stretching, symmetric and asymmetric. There are many types of bends, Twisting, Rocking, Scissoring, Tortional, Breathing (For ring molecules) and other specialized bends. Symmetric stretches require lower energy than asymmetric stretches. For a simple stretching vibration,

see: http://chipo.chem.uic.edu/web1/ocol/spec/IR1.htm p where k is a spring constant for the tensile deformation of a bond (bond strength) and u is the geometric mean mass of the two atoms, u = m1 m2/(m1 + m2) at the ends of the bond. This is similar to Campbell and White's description of LAM modes in Raman for polymer crystals which we will discuss near the end of this chapter,

Campbell and White pp. 77 and R. G. Snyder, S. J. Krause, J. R. Scherer, J. Pollym. Sci. Polym. Phys. ED. 16 1593 (1978, where n = 1, 3, 5, 7, 9... where L is the length of a sequence of C-C bonds, n is the order of the vibration (like the modes of a guitar string), [rho] is the density and E is the Young's modulus for deformation of a series of C-C bonds. These equations quantify that higher mass leads to smaller frequencies and stronger bonds lead to higher frequencies of vibration. These equations also indicate that IR absorption is a critical tool for the quantitative determination of certain molecular features such as bond strength or bond modulus. In most IR books one considers simple molecules first to gain a feel for the position of absorption peaks in the IR spectrum. Usually water and carbon dioxide are discussed first. Often both water and CO2 are present as impurities in polymers so it is important to be able to identify these bands which are not related to the material. Both of these molecules have 3 atoms. Water is not a linear molecule so 3n - 6 = 3 vibrations are expected in the IR and Raman spectra. The three vibrations are shown below (movies of these vibrations are available from the internet site mentioned).

Symmetric Stretch Asymmetric Stretch Symmetric Bend 3652 cm-1 3756 cm-1 712 cm-1 IR active IR active IR active

From: http://chipo.chem.uic.edu/web1/ocol/spec/IR1.htm (see main site above). The three vibrations are symmetric and asymmetric stretching of the H (white ball) - C (red ball) bonds, and a symmetric bending (scissors bend) of the H-C bonds. There is no asymmetric bend since such a vibration would result in a rotation of the molecule and no relative change of the position of the atoms. These vibrations follow the general rule that there are N-1 stretches (3-1=2) and 2N-5 bends (6-5=1). The symmetric stretch is an easier deformation than the asymmetric stretch so the asymmetric stretch occurs at a higher wavenumber. The bending vibration is much easier than stretching so this occurs at a much lower wavenumber. Carbon dioxide also displays, 2 stretching and 1 bend vibration.

Symmetric Stretch Asymmetric Stretch Symmetric Bend 1340 cm-1 2350 cm-1 666 cm-1 Raman active IR active IR active Not IR active Chloroform and deutero-chloroform are a common example of the effect of mass on absorption bands.

Chloroform deutero-Chloroform H-C or D-C stretch 3035 cm-1 2250 cm-1 Other Band (Deutero Bend) 1224 cm-1 910 cm-1 Cl Umbrella Bend 765 cm-1 740 cm-1 Group Contribution Method: As noted above, the large number of atoms in a polymer chain makes the possible number of IR/Raman bands enormous, 3n-6, where n is on the order of 5,000 to 10,000. Synthetic polymer chains are composed of repeated chemical groups, mer units, which are arranged about the chain axis in a similar fashion for all of these groups. The simplest approach to considering the IR/Raman absorption patterns from synthetic polymers is to identify characteristic chemical groups which give rise to absorptions. This approach, of identifying chemical groups as independent contributions to a complex IR/Raman pattern, is called the group contribution method. The basic assumption of the group contribution method is that vibrations from most chemical species are little effected by their bonding to the polymer chain. This approach is accurate in the sense that absorptions for most chemical groups will fall in a limited range which can be distinguished from other absorptions due to the strength of the absorption, for example polar bonds have strong absorptions in IR, combined with the range of wavenumber where the absorptions occur. Below are three "cheat sheets" for IR group contributions which you should be familiar with. http://chipo.chem.uic.edu/web1/ocol/spec/IRTable.htm (See main site above).

This chart and the following table are from: From: http://chipo.chem.uic.edu/web1/ocol/spec/IRTable.htm (see main site above). * 3700 - 2500 cm-1: X-H stretching (X = C, N, O, S) * 2300 - 2000 cm-1: CX stretching (X = C or N) * 1900 - 1500 cm-1: CX stretching (X = C, N, O) * 1300 - 800 cm-1: C-X stretching (X = C, N, O) http://chipo.chem.uic.edu/web1/ocol/spec/IRTable.htm (See main site above).

This chart and the following table are From a course on polymer chemistry by D. Tirrell and T. J. McCarthy.

C-H Stretch Most polymers contain C-H bonds and this is a fairly polar bond so that a strong IR absorption band is usually observed for the C-H stretch near 3000cm-1. The Following two figures and text are from Paul R. Young's web page, The 4 examples serve to demonstrate the importance of the C-H stretch in identification of organic materials such as polymers. http://chipo.chem.uic.edu/web1/ocol/spec/IR1.htm

"The infrared spectrum of benzyl alcohol displays a broad, hydrogen-bonded OH stretching band in the region 3400 cm, a sharp unsaturated (sp) CH stretch at about 3010 cm and a saturated (sp) CH stretch at about 2900 cm; these bands are typical for alcohols and for aromatic compounds containing some saturated carbon. Acetylene (ethyne) displays a typical terminal alkyne CH stretch, as shown in the second panel." Methylene, -(CH2)-: The spectra of nylon 6-6 above has an example of methylene CH stretch. This is a doublet to the right of 3000cm-1 (lower wavenumber) 2926 (a) 2853(s). A scissors bend is also seen at about 1465cm-1. Methyl, -(CH3): Methyl groups (such as in polypropylene, {-(CH2)-(CH(CH3))-}) also display a doublet in the CH stretch region just below 3000cm-1, 2962 (a), 2872 (s). The bend vibration for methyl groups is a doublet (methylene is a singlet), 1450 (a) and 1375 (s). The symmetric bend for a methyl group is called an umbrella bend vibration for obvious reasons. Compare PE, PP and polyisobutylene IR spectra below. You should be able to distinguish these three spectra. Notice the broadening of the CH stretch region due to both methyl and methylene groups in PP an PIB. Near and between 1400 and 1500 PE has a single peak for the CH bend. For PP and PIB this methylene bend is present also. Below 1400 a doublet appears for PP and PIB indicating methyl bends in asymmetric and symmetric modes. Other features in the fingerprint region (below 1500) are distinctive for the 3 polymers.

Alkenes, C=CHR: Unsaturation has a signature effect on the CH stretch shifting it to higher wavenumbers, 3020-3080 (to the left of 3000). It is easy to identify unsaturated hydrocarbons in IR by the CH stretch region. The figure below is an IR transmission spectra for polystyrene -(CH2)-(CH(C6H5))-. The aromatic ring, (C6H5), gives rise to the group of bands above 3010. The bands below 3010 are from the saturated main chain CH groups. Aromatics display a distinctive C=C stretch at about 1600cm-1 which in combination with unsaturated CH stretch above 3000 identifies polymers containing aromatics. The ring breathing vibration at 1600 is always very sharp and strong.

Polyisoprene displays a broad CH stretch region associated with a mixture of saturated and unsaturated CH stretches, the broad band at

about 960 and the sharp band at 750 are associated with out of plane bends for the CH bond attached to the C=C bond. In simple alkenes these two bands are used to distinguish between trans and cis stereoisomers, 960 for trans (H on opposite sides of the C=C bond) and 750 for cis (H on same side of C=C bond).

Alkynes, C---C-H (triple bond): As the CC bond becomes stronger the CH stretch vibration goes to higher wave number (3300cm-1 for alkynes). A weak resonance at 2100 to 2200 for the triple bond occurs. The latter bond vibration has a low dipole change so is weak in IR but strong in Raman. Carbonyl, C=O: The C=O stretch is the most distinctive absorption in IR due to the high change in dipole moment on vibration and the unique range of wavenumber where this vibration occurs, 1700 to 1780cm-1. The carbonyl stretch occurs in many commodity polymers such as polycarbonate, polyvinylacetate, polymethylacrylate and in nylon (see above, IR Raman comparison). For acetone the C=O stretch occurs at 1724 (CH2O), for aldehydes (R-CH=O) at 1730 and for methyl acetates at 1745cm-1 (R-(C=O)-OCH3). For methyl acetates the O-CH3 stretch occurs at 1100 to 1280cm-1. You should be able to identify the carbonyl contribution to the following spectra.

Alcohols, (-OH): The OH stretch occurs at 3500 to 3650 cm-1 (higher than CH). A C-O stretch occurs at 1100 to 1200 cm-1. These can be identified in the spectra of polyvinyl alcohol below.

Amines, (-NH): The N-H stretch occurs between that of CH and OH. It is also intermediate in the strength of the IR absorption, see nylon Raman and IR patterns above. Nitriles, (-C---N) triple bond CN: The CN is a strong absorption band which is seen in polyacrylonitrile (PAN) and copolymers with styrene (styrene acrylonitrile copolymers SAN). The absorption occurs in the 2200 to 2300 cm-1 range.

Halides, (C-Cl, C-F): The C-Cl stretch is a strong absorption at low wavenumber, 760-540 cm-1. Two examples of halide stretch from PVC (CH2-CHCl)- and teflon -(CF2-CF2)- are shown below.

Other Examples from Web: "Saturated and unsaturated CH bands also shown clearly in the spectrum of vinyl acetate (ethenyl ethanoate). This compound also shows a typical ester carbonyl at 1700 cm and a nice example of a carbon-carbon double bond stretch at about 1500 cm. Both of these bands are shifted to slightly lower wave numbers than are typically observed (by about 50 cm) by conjugation involving the vinyl ester group."

Selected Applications of IR/Raman: Tacticity and Crystallinity: In some cases it is possible to assign certain absorption bands with tacticity in polymer chains. In most cases only a qualitative measure of tacticity is gained from IR and Raman spectroscopy. Figures 5.13 and 5.18 of Campbell and White and the following figure show the qualitative differences observed in IR/Raman spectra for tactic forms of polypropylene. In some cases specific bands are associated with specific conformations which are possible in tactic forms. Many of these conformational differences disappear when samples are run in the melt. There is a significant overlap between tacticity and crystallinity determinations in IR analysis of polymers.

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993. Raman spectra form polypropylene particles. Atactic: weak, diffuse Isotactic: sharp and strong. CH3 umbrella bend is to the left. The presence of crystallinity also leads to predictable changes in IR patterns and has been used to qualitatively determine the degree of crystallinity for instance.

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993. LAM Modes in Raman (Crystallite Thickness): Low frequency regions (< 30cm-1 in polyethylene) have been associated with longitudinal acoustical modes (accordion modes) for the planar zigzag chain conformation in a lamellar crystallite. From the frequency of absorption, the length of such a planar zigzag chain can be determined using a simple model:

From R. G. Snyder, S. J. Krause, J. R. Scherer, J. Pollym. Sci. Polym. Phys. ED. 16 1593 (1978, where n = 1, 3, 5, 7, 9...

From R. G. Snyder, S. J. Krause, J. R. Scherer, J. Pollym. Sci. Polym. Phys. ED. 16 1593 (1978).

From R. G. Snyder, S. J. Krause, J. R. Scherer, J. Polym. Sci.; Polym. Phys. Ed. 16 1593 (1978). The figure above shows a comparision between the LAM method and the use of small angle x-ray scattering (SAXS) to determing the lamellar thickness. From such a comparison the constants in the LAM equation can be determined. For PE the lamellar thickness, L ~ 6000/[nu]. Orientation: If polarized radiation is used in IR/Raman it is possible to determine the relative orientation of specific absorbing groups in a processed sample. The usual way to do this is to determine the Herman's Orientation function, f, for the group of interest. f has a value of 0 for unoriented samples, 1 for samples oriented in the machine direction and -1/2 for samples oriented perpendicular to the machine direction but in the plane of observation. The absorption ratio for a given bond is measured by rotating the sample parallel and perpendicular to the incident direction of polarization, R = Aparallel/Aperpendicular. The angle between the bond axis and the polymer chain axis, a, needs to be determined from molecular models if chain orientation is of interest. The orientation function is then given by:

The first function is a generic description of the uniaxial Herman's orientation function. The second function is calculated for directional absorption in IR. The following figure from Hunt and James shows the complexity involved in calculation of IR absorption orientation. Wilkes has written a good review article on this subject, G. Wilkes (1971), Adv. Polym. Sci., 8, 91.

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993. The IR and Raman (lower) patterns below show the type of changes in IR and Raman patterns which are observed for oriented samples of PET.

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993. The orientation function can be calculated through a number of techniques some of which are shown in the following figure for PET as a function of draw ratio.

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993. Real time studies of deformation can yield information as to which chemical groups are involved in mechanical manipulation of samples. The following spectra are from continuous deformation of a polyether-polyurethane sample. These show the wealth of information which is available in a rheo-optical study using IR/Raman.

From "Polymer Characterization" by B. J. Hunt and M. I. James, Blackie Press 1993.

= Back to TOC = To Syllabus

Chapter 6. NMR Spectroscopy (Chapter 6 Campbell & White). http://www.shu.ac.uk/schools/sci/chem/tutorials/molspec/nmr1.htm http://www.informatik.uni-frankfurt.de/~garrit/biowelt/nmr.html http://www-wilson.ucsd.edu/education/spectroscopy/nmr.html (Physics) Introduction: Nuclear magnetic resonance is an absorption spectroscopy involving the absorption of radio frequency EM waves. Since we have already covered IR absorption spectroscopy it is appropriate to compare these two techniques, building on what we already know. The energy associated with a photon in the radio frequencies is extremely small compared to IR frequencies. In fact, we are constantly being irradiated by radiowaves with no effect. Absorption spectroscopies rely on the transfer of energy (h) from an electro-magnetic wave to a quantized transition in a material. For IR the quantized transition (a transition with a fixed energy) is the vibration of a chemical bond. NMR involves changes in the spin state of the nucleus of an atom. Not all nuclei display spin. In order to display spin a nucleus must have an odd number of protons or neutrons. Hydrogen has one proton, so displays spin (1H or proton NMR). Deuterium has one proton and one neutron and also displays spin. Other atoms that display spin are isotopes of common elements (deuterium is an isotope of hydrogen). The most common are 13C, 19F, 15N and 29Si. An atom with spin has a non-zero spin quantum number , I. For most nuclei of interest to polymer scientists the spin quantum number is 1/2. Deuterium and Nitrogen 15 have spin quantum numbers of 1. The number of spin states possible for a nucleus is given by 2I+1. For most nuclei of interest to polymer scientists there are two spin states. IR absorption also involves two states, i.e. vibrating and not vibrating. In NMR the two states correspond with two orientations of magnetic moment vector for the nucleus with respect to an external magnetic field as discussed below. For a quantized transition to occur a system must be constrained, i.e. a guitar string must be under tension for the production of a note, atoms must be bonded for an IR absorption. In NMR the constraint which leads to quantized transitions is applied to the sample externally in a large static magnetic field. If a nucleus spins and is composed of charged particles it possesses a magnetic moment associated with the angular velocity of the charged particles. The magnitude of the magnetic moment, , is proportional to the spin quantum number,

= gN N I
gN is a constant and N is the nuclear magneton, given by,

N =

eh 4mc

where all the parameters are constant and related to a proton. The magnetic moment is a vector, , and in vector form it is defined as,

=I
where I is the angular momentum vector, given by. I= h 2 I ( I + 1)
1

and is the magnetogyric ratio for the nucleus. The frequency of absorption in IR is directly proportional to the magnetiogyric ratio,

= B0 =

2B0 Ih

where B0 is a strong magnetic field which is applied to the sample. The nucleus can then be thought of as a magnet, and in the absence of a magnetic field these tiny magnets are randomly arranged in a sample with no preferred direction for the magnetic moment vectors. Application of a radiofrequency EM wave to such a sample has no effect, i.e. there is no absorption. There is not absorption because the system can no tell the difference between magnetic vectors pointing up or down or any other direction since these directions have not reference base. This is analogous to a guitar string which is not constrained o r two atoms which are not bonded in IR. In the absence of constraints there is n o perceptible absorption. If a strong magnetic field is applied to the nuclei, they can tell the difference between alignment in the direction of the applied magnetic field and opposed to the applied magnetic field. In NMR the constraint which leads to quantized transitions is applied by the spectrometer. Because of this the frequency of absorption varies with the applied magnetic field and there is no absolute frequency or wavelength for a given absorption. NMR spectra are not plotted as absorption versus wavenumber as IR spectra are, they are plotted as absorption versus chemical shift, . The chemical shift for proton NMR is the difference between the frequency of absorption of the sample and a standard, tetramethylsilane (TMS) normalized by the frequency of absorption of TMS,

Sample TMS TMS

*10 6

is expressed in parts per million (ppm) so the above equation is multiplied by 106. In IR we consider two states for a bond, vibrating and non-vibrating. The transition associated with the change from non-vibrating to vibrating leads to the absorption at fixed wavenumbers. In NMR several states are potentially possible depending on the spin quantum number. The permitted states are given by the allowable values for the magnetic quantum number, mI = I, I-1,...-I. For I = 1/2 there are two states possible, mI = 1/2 and mI = -1/2. For I = 1 three states are possible, mI = 1, 0, -1. For protons the two allowable states are generally spoken of as parallel and anti-parallel to the applied field. Intensity of Absorption: The strength of an IR absorption band depends on the change in dipole moment for the bond on vibration, i.e. how polar a bond is. The strength of a NMR absorption band depends o n the magnitude of the magnetogyric ration, , i.e. how large the magnetic dipole moment is. The absorption in IR is also proportional to the concentration of the absorbing bond. NMR depends on the presence of specific isotopes. In considering the strength of a NMR absorption band we consider the "natural abundance" of these isotopes. For example, 99.98 percent of hydrogen atoms are 1H, and 0.0156 percent are deuterium, 2H. The magnetogyric ratio for hydrogen is 26,700 while for deuterium is 4,100. This means that proton NMR (1H) results in
2

100 times the signal as deuterium if a sample contains natural abundance hydrogen. A similar comparison shows that proton NMR absorption is about 50 times stronger than 13C NMR (natural abundance of 13C is about 1%). An NMR absorption peak for a given nucleus is directly proportional to the number of these atoms in a sample. A major difference between analysis of NMR spectroscopy in polymers and IR spectra is that all absorption bands in NMR are uniquely identifiable. The NMR Experiment, Pulsed NMR: The simplest NMR experiment is the observation of "free induction decay" from a radio frequency pulse. This is analogous to the plucking of a guitar string. The free induction decay looks similar to the raw data obtained form a FTIR instrument except that the x-axis is time rather than space. Fourier transform of the spectra results in inverse units, i.e. frequency, which is converted to .

(1 )
Precession

(2 ) N RF Pulse 90 to B0
S N

(3 ) N
S N

N
S N

Observe Free Induction Decay

t0

td

RF Signal

/2 Free Induction Decay Pulse

Time
In the NMR experiment many pulses are applied and the FID patterns summed to enhance the signal. The delay time between pulses, td, determines the smallest frequency which is observed. The width of the initial /2 pulse, t 0, determines the largest frequency which is observed. The summed decay pattern is Fourier transformed to obtain the spectrum in frequency.

The NMR instrument is capable of many other more complicated experiments involving sequencing of pulses and observation of kinetic phenomena in spin relaxation which will be discussed later in this chapter. The nuclear dipole tilts at an angle with respect to the static magnetic field when the RF pulse is applied. The magnetic dipole then rotates about the static field at this precession angle, . The precession angle is determined by the RF field strength, H 1, the pulse time, t0, and the magnetogyric ration of the nucleus, .

= H1t0
Position of Absorption Peaks in NMR Spectra: Consider an isolated nucleus in a static magnetic field of B0 = 14,000 Gauss. The frequency of absorption for different nuclei varies according to the magnetogyric ratio,

14 N

2H

13 C

31 P

19 F

1H

8 13 58 60 Frequency in Mega Hertz (MHz)

This would be called a 60 megahertz NMR instrument because the proton NMR resonance is close to 60 megahertz. A typical proton NMR spectra will go from 0 to 10 ppm meaning that the entire spectra will be 60MHz 0.0006 MHz. The absorption from other nuclei will not effect the proton spectra at all since their absorptions are at far different frequencies. NMR is extremely sensitive to the "chemical environment" of a nucleus. "Chemical environment" means the local magnetic environment. The local magnetic environment is changed by "shielding" or "deshielding" depending on how the chemical bonds which are attached to the nucleus withdraw electrons from the electron cloud of a bare atom. Electron withdrawing groups such as the aromatic ring, deshield a proton for instance and give rise to a deshielded proton with a large . Tetramethyl silane (TMS) has highly shielded protons so is used as a standard for the 0 point of a NMR spectra. The only absolute point for a given nucleus would be a nucleus stripped of all electrons. Since it is not possible to obtain such a completely deshielded nucleus TMS is used as a standard.

Phenet hyl Iodid e


H H H H I H H H H H

-10

ppm

Shielded Deshielded

0 TMS

The chemical environment of protons changes the frequency of absorption as shown in the cartoon of the absorption spectra for phenethyl iodide above. TMS is highly shielded so is a good 0 point for frequency. The "Chemical Environment" reflects the bonding geometry of a molecule. Two protons with identical chemical environments "see" the same view of the molecule from their position in the bonding structure. The same view includes the stereochemical arrangement of the molecule, i.e. the handedness of the structure. The "view" which effects proton absorption is three bonds in distance, that is, structural differences more than three bonds away don't matter. Additionally, the presence of resonance structures such as the aromatic ring in phenethyl iodide makes all of the aromatic protons see the same chemical environment. The aromatic protons in phenethyl iodide are identical. The methyl iodide group at the other end of phenethyl iodide has two protons with identical chemical environments, so give rise to a single peak at the far right of the spectrum. The two methylene protons have different chemical environments because of their stereochemical relationship to the methyl iodide group. Protons with the same chemical environment give rise to a single peak, so simple or symmetric molecules have single peaks in proton NMR spectra. Examples are water, benzene and cyclohexane. The aromatic ring in benzene is strongly electron withdrawing, so the protons are strongly deshielded. The oxygen in water is also electron withdrawing and deshielding. The cyclohexane ring is only weakly electron withdrawing so weakly deshielding in proton NMR.

H H C H

H C H

O
H H H H Cyclohexane

H
C H

Si
C

H H H

Benzene -7. 3
H H H

Water -5. 2

-1. 4
H H

H H

0 TMS
H H H

H H H H H H H H

The intensity in proton NMR is directly proportional to the number of protons in the structure with identical chemical environments. Toluene has 5 aromatic protons (-7.3ppm) and 3 methyl protons (-2.5 ppm) with an absorption intensity ratio of 5:3.

H C H H

H H

Toluene -7. 3 -2. 5 0

Multiplet Splittings: Nuclei are magnetically coupled through 3 or fewer bonds. Since each proton can exist in one of two states, I = 1/2 or I = -1/2, neighboring protons (3 or fewer bonds away) will create 2 magnetic environments for a given proton. A single absorption peak will be split into two peaks for a single neighboring proton. The number of splittings is given by n+1 where n is the number of neighboring (within 3 bonds) magnetically equivalent protons. If there are two types of neighboring (within 3 bonds) protons the number of splittings multiplies. (The general rule is 2nI + 1 splittings.) These multiplet splittings will be centered on the normal absorption frequency or of the proton. For example, in isopropyl benzene the 6 methyl proton band will be split into two bands of equal intensity by the single methylene proton. The 6 methylene protons will split the single methylene proton into 2nI+1 absorption bands or 7 bands (I = 1/2). Definition of Magnetic Equivalence: Protons are equivalent if: 1) Same chemical shift, . 2) Same coupling constant to all nuclei in the molecule, JHa. Examples: (CH3)2CHBr 2 types CH3 and CH Propene CH3CHCH2
H3 C Hcis C H C Htrans

4 types

H cis, Htrans, CH3, CH

Magnetic Equivalence Means Same Magnetic Environment There is not way to magnetically distinguish the nuclei The Nuclei's view of the molecule is the same both in terms of bonds and in terms of stereochemistry. Nuclei can only sense 3 bonds away, with no conjugated double or triple bonds between them and can only sense other nuclei with similar nuclear magnetic transitions.

H H H H C C

H H H C H

H H H H

Iso-Propyl Benzene

-7. 3

-3

-1

The frequency of the distance between splittings of an absorption band i s independent of the applied field. This is called the J coupling constant. The J spin coupling constant is an absolute value for a given pair of nuclei, i.e. JHMethyl is a fixed value on a frequency scale (but not on a scale). The intensity of the split bands follows the binomial distribution (n is the number of chemically identical protons splitting a given band): n 1 2 3 4 5 6 Relative Intensity of split bands. 1 1 1 2 1 1 3 3 1 1 4 6 4 1 1 5 10 10 5 1 6 15 20 15

1 6

For coupling by non-chemically identical protons the binomial distributions multiply, i.e. for splitting by two non-chemically identical protons 4 bands of equal intensity would result. Unlike IR, all absorption bands in an NMR spectrum can be completely identified. This makes NMR an extremely powerful tool for chemical identification. NMR is also extremely flexible since you have complete control over the constraint which leads to the absorption. The drawback to NMR is the cost of the instrument, $250,000 to $1e6, and the expertise needed for more than routine identification. The cost and expertise are roughly both an order of magnitude higher than IR. Example: Ethanol

Chemical Shift Series: Shielding is directly related to the electron density around the nucleus. There are a number of homologous series (series varying a chemical group to observe the chemical shift) which demonstrate this.
H H H

+
H

C H

e+ Chemical Shifts/Electronegativity for CH 3 C H 2 X X Electronegativity( ) Chemical Shift for CH2 -SiEt3 1.9 0.6 -H 2.2 0.75 -CEt3 2.5 1.3 -NEt2 3.0 2.4 -OEt 3.5 3.3 -F 4.0 4.0 Double bonds and Rings with conjugated bonds (high electron orbitals) deshield associated protons: Compound ethane ethylene ethyne Structure CH3CH3 CH 2=CH2 HC=CH Chemical Shift for CH2 0.9 5.0 (Current) 2.3 (No Current)
9

Benzene

-H

7.3

(Current)

Parts of the NMR Spectra: In summary the NMR spectra is composed of absorption peaks at a value of chemical shift, , relative to a reference material, usually TMS. The amplitude is proportional to the number of magnetically equivalent protons of that type in the molecule. When neighboring protons (3 bonds or less away) are present the peak amplitude is split into several peaks following the binomial distribution and separated by the J coupling constant which is fixed in value, in terms of frequency, of the J coupling (not ) according to the type of proton which is causing the splitting (see figure below):

From Hunt and James, "Polymer Characterization" Tacticity: The tetrahedral bond of carbons can be thought of as a tripod with a bond sticking straight up. If a polymer chain is attached to the bond sticking straight up and the chain is attached to one of the legs of the tripod then substitutent groups attached to the other two legs have a choice of being placed to the right or left. This can be depicted in a Neuman projection along the chain back bone where the circle and three line apex are two carbons along the main chain connected by a bond.

10

Polymer Chain

Ha

Hb

H Polymer Chain

Cl

The presence of the Cl substitutent group allows for the distinction of Ha and H b. H a is gauche to Cl and Hb is anti to the Cl group. For each mer unit in this polyvinylchloride molecule there will be an anti and a gauche methylene proton. The relationship between these mer unit stereo chemistries gives rise to tacticity . Because of this the smallest grouping for tacticity is a diad, two mer units. Diad Tacticity: For two substitutent groups in a three carbon sequence the substitutents can be located with the same handedness (Meso) or with opposite handedness (Racemic).
Cl H Cl C C Ha Hb H

Meso a and b not equivalent Anti to Cl's

Gauche to Cl's
Cl H H C C Ha Hb

Cl

Racemic a and b equivalent

Both are Gauche to 1 and Anti to other


11

From an NMR perspective, Racemic diads give rise to two magnetically equivalent protons on the methylene group while Meso diads give rise to two magnetically different protons on the methylene group. NMR can not sense diad tacticity because the proton on the substituted carbon can sense two diads (on either side). The smallest unit of tacticity which NMR can detect in polymers is a triad (3 mer units). Because of this triad tacticity is the usual way to refer to polymer stereochemistry. NMR can also sense higher o d d number groupings of tactic mer units, with diminishing resolution, pentads ( 5 ) , heptads (7) etc.

H R C C

H R C C

H R C C

If H "sees" this mer

H also "sees"
this mer

Note that the handedness, anti or gauche, is independent of rotation about the C-C bonds which occurs in all single bond chains. The particular arrangement shown above is merely for convenience of comparison between meso and racemic diads and does not reflect the actual conformation of the chains in a polymer which would be reflected by a distribution of rotational orientations reflecting the potential energy diagram for C-C bond rotation as show below for polyethylene (PE does not display tacticity),

C H H

H C

Pot ent ial Energy

H H H

H H

Rot ati on
Triad Tacticity: A triad is composed of two diads which share a central mer unit. There are three possibilities for triad tacticity based on diad tacticity: Isotactic, meso+meso Syndiotactic, racemic+ racemic Heterotactic, racemic + meso or meso + racemic
12

mm rr rm or mr

1 1 2

A polymer with no preferred tacticity, an atactic polymer, has a random statistical distribution of diad tacticities so it would have 25% isotactic triads, 25% syndiotactic triads and 50% heterotactic triads. A polymer with 50% meso and 50% racemic diads does not necessarily have an atactic triad distribution, just as an atactic triad distribution does not necessarily have an atactic (random) pentad distribution. An atactic (random) pentad distribution does imply atactic triad and diad distributions. Since syndiotactic is composed of two racemic units, and because the protons in a racemic diad are magnetically equivalent (see above), then syndiotactic triads will have the fewest number of magnetic types of protons and the fewest peaks and splittings. This is shown for PMMA in figure 6.6 of Campbell and White shown below (isotactic top, syndiotactic bottom): -COOCH3 -CH 3 -CH 2
H H C H

() -COOCH3 -CH 3 -CH 2


Polymer C H O

()
C C O C H H H Polymer

PMMA

From Campbell and White "Polymer Characterization" PMMA was one of the first polymers studied in depth for tacticity using proton NMR (see texts by Bovey from the 1970's). For syndiotactic PMMA (bottom) 3 main absorptions are observed, CH3 at 0.91, -CH2 at 1.9 and -COOCH3 at 3.6. For isotactic polymer (top curve) the -CH3 is more deshielded 1.20, the -CH2 becomes a quartet centered at 1.9 and the -COOCH3 remains a singlet at 3.6. The splittings of the -CH2 in what should be a sequence of 1:1:1:1 is due to two types of methylene groups, termed erythro, e, (more deshielded) and threo, t, (less deshielded) corresponding to the bottom and top protons in the molecular sketch above. Each of these peaks are split into two peaks by the other leading to and expected splitting of 4 equal peaks, with a reported J coupling constant of about 0.2 ppm. The e and t protons are separated by 0.7 ppm which can be verified by molecular modeling. A higher resolution NMR can resolve higher order stereosequences as shown below for isotactic and atactic PMMA. You should compare the information content of the 60 MHz spectrum above to the 500MHz spectra below. (Again, 60 MHz refers to the natural resonance frequency of a proton for a given magnetic field of the instrument, B0.)
13

From Hunt and James "Polymer Characterization"

14

From Hunt and James "Polymer Characterization" A similar comparison of signal can be made for the 100 and 500 MHz spectra of polyvinyl chloride given below. -methylene protons occur at 1.5-2.5 range and the single -methine proton is observed at about 4.6. In many polymers resolution of splittings for individual stereochemical peaks is not possible and this is illustrated by the PVC spectra. In such cases the tactic sequences can be identified with complicated NMR techniques such as 2-D NMR (see Campbell and White). Generally you will find a reference which has identified the peaks associated with certain tactic groupings for vinyl polymers and use the integrated areas of these peaks to determine the triad tacticity of a given polymer. (Note that the PMMA example given above is the best resolved spectra of commodity polymers, i.e. sharpest peaks, and this is related to the structure of PMMA). Methine
H Polymer H

( )

C H

()

C Cl

Polymer

Methylene

15

From Campbell and White, "Polymer Characterization"

From Hunt and James "Polymer Characterization" The splittings of the tactic peaks in the proton NMR spectrum of PVC, shown above, are not resolvable on typical NMR spectrometers. Use of a different nucleus, 13C, can overcome problems with resolution of this type. The 125 MHz, 13C spectra for PVC is shown below. Notice the higher resolution even compared to the 500 MHz proton spectra shown above. (Note that spectrometer magnetic field strength is in reference to the proton resonance frequency even if a different nuclei is probed.)

From Hunt and James "Polymer Characterization"


13

C NMR:
16

1) Low Natural Abundance: Since most polymers are composed of hydrogen and carbon, the natural alternative nucleus for NMR is 13C. There are a number of major differences between proton and carbon 13 NMR. First, the natural abundance of 13C is much lower than 1H (12C does not display spin since the number of protons and neutrons are both even). The natural abundance of 13C is about 1.1 % while that of 1H is close to 100%. Since only nuclei of similar magnetic resonance can lead to coupling and splitting of the absorption peaks, the low natural abundance of 13 C leads to no splittings of the absorption peaks. The sensitivity of absorption of a RF pulse and the associated decay are also much lower for 13C. 2) Large Chemical Shifts: The range of proton absorptions are on the order of 10ppm relative to TMS. For 13C the range of absorptions are on the order of 200ppm relative to TMS. The 13C spectrum has more than an order higher resolution when compared to 1H spectra as can be seen in the PVC spectra above. 3) The large abundance of 1H nuclei compared with 13C leads to loss of 13C resolution and signal due to weak coupling of 13C and 1H resonances. This problem is amplified in solid samples, so called solid state 13C NMR. Cross Polarization: The low abundance of 13C leads to poor absorption of the RF pulse in a FT-NMR experiment. This limitation can be over come by exciting the protons in a sample followed by a sequence of two series of long-time pulses which make the 13C and 1H nuclei resonate at the same frequency. The latter is called the "Hartman-Hahn" condition and the process is called "cross-polarization" and the time of cross polarization is called the "contact time" or "spin-lock time". This cross polarization acts as a strong pulse for the carbon 13 nuclei (see figure below).

From Hunt and James "Polymer Characterization" Proton Decoupling: Cross-polarization leads to a large enhancement of the excitation of 13C nuclei. The large number of 1H in the sample, however, interfere with the decay of the isolated 13C nuclei due to weak interaction of the spins. For example, this would be like trying to play a guitar under water, that is
17

despite the difference in resonance frequency for a swimming pool full of water and the guitar string, there is transfer of energy to the pool from the guitar. This dampening of the 13C signal can be removed by a strong radio frequency signal which essentially holds the protons in a highly resonating state so they are not capable of absorbing resonance from 13C nuclei. Cross polarization and spin decoupling were critical developments for the wide use of 13C NMR. The figure below shows the effect of proton decoupling on a carbon 13 NMR signal.

From Hunt and James "Polymer Characterization" Magic Angle Spinning and Solid State 13 C NMR: All of the previous discussion was based on "solution NMR" where a polymer sample is dissolved in a solvent at 1 to 20% concentration. One reason for studying polymers in solution is that the anisotropy of the magnetic moment with respect to the macromolecule (Chemical Shift Anisotropy, CSA) is averaged out due to thermal motion of the molecule in solution. Chemical shift anisotropy has the effect of smearing out the NMR signal as can be seen by comparison of the bottom and second to the bottom spectra in the previous figure. In the solid state, i.e. a semicrystalline or glassy polymer, CSA has a severe effect on the spectra in broadening the absorption peaks and the effect becomes worse the higher the restriction of mobility of the chains or molecules. Through a tensoral analysis of the magnetic moments in a molecule it is possible to demonstrate that a "Magic Angle" exists with respect to the applied magnetic field at which rapid
18

spinning of the solid sample leads to minimization of absorption line broadening due to chemical shift anisotropy. Examples of 13 C NMR spectra: Several examples of 13C NMR spectra from Campbell and White, and Hunt and James are given below. The PMMA spectra should be compared with the proton NMR spectra given above.

From Campbell and White "Polymer Characterization"

From Campbell and White "Polymer Characterization"

From Campbell and White "Polymer Characterization"


19

From Hunt and James "Polymer Characterization" Other Nuclei: A number of plastics and elastomers are based on nuclei other than 13C or 1H. Two examples are given below, 29Si (natural abundance 4.7%) and 19F (natural abundance 100.0%). Silicon 29 NMR is conducted similar to carbon 13 NMR while Fluorine 19 parallels closely proton NMR. The advantage of using these alternative nuclei is that the degree of C substitution on Si can be directly determined and a much higher resolution of fluorinated tacticity can be determined.

20

From Hunt and James "Polymer Characterization"

From Hunt and James "Polymer Characterization"


21

Chapter 7. XRD (Chapter 8 Campbell & White, Alexander "X-ray Diffraction Methods in Polymer Science"). The general principles of diffraction are covered in Cullity, "Elements of X-ray Diffraction". If you are unfamiliar with XRD you will need to review or read Cullity Chapters 1-7 and the appendices. Alexander's text referenced above is also useful as an introduction to XRD but is less general and at a slightly more advanced level. There are a number of differences between x-ray diffraction in polymers and metallurgical (Cullity) or ceramic diffraction. 1) Polymers are not highly absorbing to x-rays. The dominant experiment is a transmission experiment where the x-ray beam passes through the sample. This greatly simplifies analysis of diffraction spectra for polymers but requires somewhat specialized diffractometer from those commonly used for metallurgy (usually a reflection experiment).

From Alexander, "X-Ray Diffraction Methods in Polymer Science" For transmission geometry the optimal sample thickness is 1/ where is the linear absorption coefficient. Typically the optimal thickness for a hydrocarbon polymer is 2 mm. (See Cullity for calculation of optimal thickness for a diffraction sample in transmission). 2) DOC: Polymers are never 100% crystalline. XRD is a primary technique to determine the degree of crystallinity in polymers. 3) Synthetic polymers almost never occur as single crystals. The diffraction pattern from polymers is almost always either a "powder" pattern (polycrystalline) or a fiber pattern (oriented polycrystalline). (Electron diffraction in a TEM is an exception to this rule in some cases.)
1

4) Microstructure: Crystallite size in polymers is usually on the nano-scale in the thickness direction. The size of crystallites can be determined using variants of the Scherrer equation. 5) Orientation: Polymers, due to their long chain structure, are highly susceptible to orientation. XRD is a primary tool for the determination of crystalline orientation through the Hermans orientation function. 6) Polymer crystals display a relatively large number of defects in some cases. This leads to diffraction peak broadening (see Campbell and White or Alexander for details). 7) Polymer crystallites are very small with a large surface to volume ratio which enhances the contribution of interfacial disorganization on the diffraction pattern. 8) SAXS: Due to the nano-scale size of polymer crystallites, small-angle scattering is intense in semi-crystalline polymers and a separate field of analysis based on diffraction at angles below 6 has developed (see Alexander and Chapter 8 of these notes for details). Introduction: Diffraction or scattering is a separate category of analytic techniques using electromagnetic radiation where the interference of radiation arising from structural features is observed. The interference pattern is the Fourier transform of the pair wise correlation function. The pair wise correlation function can be constructed in a though experiment where a multiphase material is statistically described by a line throwing experiment. If lines of length "r" are thrown in to a 2 phase material there is a probability that both ends of the lines fall in the dilute phase. This probability in 3-d space changes with the size of the line, "r", and a plot of this probability as a function of "r" is a plot of the pair wise distribution function. For a crystal the two phases are atoms and voids and peaks in the pair wise correlation function occur at multiples of the lattice spacing. Interference which results from correlations of different domains or atoms is usually associated with the "Structure Factor" or "Interference Factor", S2(2). Interference can also occur if the individual domains are prefect structures such as spheres. For a sphere, there is a sharp decay in the pair wise correlation function near the diameter of the sphere and this sharp decay results in a peak in the Fourier transform of the correlation function. For a metal crystal this corresponds to the atomic form factor, f2(2 ). For larger scale domains interference associated with the form of the scattering units is generally termed the "form factor", F2(2). The scattered intensity as a function of angle is then the product of two terms, the form factor (f2(2) or F2(2)) and the structure factor (S2 (2)): I(2) = Constant F2(2) S2 (2) For XRD the form factor is usually obtained from tabulated values and the major interest is in the Structure factor. For small angle scattering dilute conditions are usually of interest making the structure factor go to a constant value of 1 and the form factor for complex structures are investigated. Thus, the basic principles of scattering and diffraction are the same, while the implementation of these principles are quite different. Bragg's Law:
2

Cullity and Alexander derive Bragg's Law using the mirror analogy (specular analogy). It can also be derived from interference laws or using "inverse space" (see appendix in Cullity). The features of Bragg's Law is that structural size is inversely proportional to a reduced scattering angle, so high angle relates to smaller structure and low angle relates to large structure. Small-angle scattering measures colloidal to nano-scale sizes. There is no large scale limit to diffraction. The small scale limit (i.e. the smallest measurable size) is /2 as is inherent in Bragg's Law: d = /2 (1/sin) is half of the scattering angle measured from the incident beam. The 1/sin term in Bragg's law acts as an amplification factor. The minimum value of which is 1 for 2 = 180 (direct back scattering). The maximum value of the amplification factor is so that theoretically no size limit exists with a given radiation of wave length . In reality the diffraction geometry and coherence length of the radiation leads to a large scale limit on the micron scale. Typically diffracted intensity if plotted as a function of 2 . Since the d-spacing is of interest one might wonder why diffraction data isn't plotted as a function of sin or 1/sin. This is in fact done with the use of the "scattering vector" q or s. q = 4/ sin() = 2/d and s = 2/ sin() = 1/d. The appendix of Cullity gives a good description of diffraction in "q" or "s" reciprocal space. The Fourier transform of the real space vector, "r", used to determine the pair wise correlation function is the scattering vector "q". Review of Crystalline Polymer Morphology: "Molecular" scale Crystalline Structure: Consider that we can form an all-trans oligmeric polyethylene sample an bring it below the crystallization temperature. The molecules will be in the minimum energy state and will be in a planar zigzag form. These molecular sheets, when viewed from end will look like a line just as viewing a rigid strip from the end will appear as a line. Crystal systems are described by lattice parameters (for review see Cullity X-ray Diffraction for instance). A unit cell consists of three size parameters, a,b,c and three angles , , . Cells are categorized into 14 Bravis Lattices which can be categorized by symmetry for instance. All unit cells fall into one of the Bravis Lattices. Typically, simple molecules and atoms form highly symmetric unit cells such as simple cubic (a=b=c, ===90) or variants such as Face Center Cubic or Body Centered Cubic. The highest density crystal is formed equivalently by FCC and Hexagonal Closest Packed (HCP) crystal structures. These are the crystal structures chosen by extremely simple systems such as colloidal crystals. Also, Proteins will usually crystallize into one of these closest packed forms. This is because the collapsed protein structure (the whole protein) crystallizes as a unit cell lattice site. In some cases it is possible to manipulate protein molecules to crystallize in lamellar crystals but this is extremely difficult. As the unit cell lattice site becomes more complicated and/or becomes capable of bonding in different ways in different directions the Bravis lattice becomes more complicated, i.e. less symmetric. This is true for oligomeric organic molecules. For example olefins (such as dodecane (n=12) and squalene (n=112)) crystallize into an orthorhombic unit cells which
3

have a, b and c different while ===90. The reason a, b and c are different is the different bonding mechanisms in the different directions. This is reflected in vastly different thermal expansion coefficients in the different directions. The orthorhombic structure of olefinic crystals is shown below. Two chains make up the unit cell lattice site (shown in bold). The direction of the planar zigzag (or helix) in a polymer crystal is always the c-axis by convention.

PE/Olefin crystal structure. See also, Campbell and White figure 8.4. Chain Folding: The planar zigzag of the olefin or PE molecule crystallize as shown above into an orthogonal unit cell. This unit cell can be termed the first or primary level of structure for the olefin crystal. Consider a metal crystal such as the FCC structure of copper. The copper atoms diffuse to the closest packed crystal planes and the crystal grows in 3-dimensions along low-index crystal faces until some kinetic feature interferes with growth. In a pure melt with low thermal quench and careful control over the growth front through removal of the growing crystal from the melt, a single crystal can be formed. Generally, for a metal crystal there is no particular limitation which would lead to asymmetric growth of the crystallite and fairly symmetric crystals result. This should be compared with the growth of helical structures such as linear oligomeric olefins, figure 4.1 on pp. 143 of Strobl. Here there is a natural limitation of growth in the c-axis direction due to finite chain length. This leads to a strongly preferred c-axis thickness for these oligomers which increases with chain length. In fact, a trace of chain length versus crystallite thickness is a jagged curve due to the differing arrangement of odd and even olefins, but the general progression is linear towards thicker crystals for longer chains until about 100 mer units where the curve plateaus out at a maximum value for a given quench depth. (Quench depth is the difference between the equilibrium melting point for a perfect crystal and the temperature at which the material is crystallized.)

Thickness

Depends on Quench Depth T

Number of mer units Schematic of olefin crystallite thickness as a function of the chain length.
The point in the curve where the crystallite thickness reaches a plateau value in molecular weight is close to the molecular weight where chains begin to entangle with each other in the melt and there is some association between these two phenomena. Also, the fact that this plateau thickness has a strong inverse quench depth dependence suggests that there is some entropic feature to this behavior (pp. 163 eqn. 4.20 where dc is the crystallite thickness and pp. 164 figure 4.18 Strobl). Considering a random model for chain structure such as shown in figure 2.5 on pp. 21 as well as the rotational isomeric state model for formation of the planar zigzag structure in PE, pp. 15 figure 2.2, it should be clear that entropy favors some bending of the rigid linear structure, and that this is allowed, with some energy penalty associated with gauche conformation of figure 2.2. Put another way, for chains of a certain length (Close to the entanglement molecular weight) there is a high-statistical probability that the chains will bend even below the crystallization temperature where the planar zigzag conformation is preferred for PE. When chains bend there is a local free energy penalty which must be paid and this can be included in a free energy balance in terms of a fold-surface energy if it is considered that these bends are locally confined to the crystallite surface as shown on pp. 161 figure 4.15; and pp. 185 figure 4.34. There are many different crystalline structures which can be formed under different processing conditions for semi-crystalline polymers (Figures 4.2- 4.7 pp. 145 to 149; figure 4.13 pp. 157; Figure 4.19, pp. 165; figure 4.21 pp. 170). As a class these variable crystalline forms have only two universal characteristics: 1) Unit cell structure as discussed above. 2) Relationship between lamellar thickness and quench depth. This means that understanding the relationship between quench depth and crystallite thickness is one of only two concrete features for polymer crystals. John Hoffman was the first to describe this relationship although his derivation of a crystallite thickness law borrowed heavily on asymmetric growth models form low molecular weight, particularly ceramic an metallurgical systems. Hoffman's law is given in equation 4.23 on pp. 166: n* = 2 m, e Tf
f Hm Tf T

Cryst allit e

Hoffman Law

where n* is the thickness of the equilibrium crystal crystallized at T (which is below the equilibrium melting point for a crystal of infinite thickness, Tf), is the excess surface free energy associated with folded chains at the lateral surface of platelet crystals, and H is the heat of fusion associated with one monomer. Hoffman's law can be obtained very quickly for a free energy balance following the "GibbsThomson Approach" (Strobl pp. 166) if on considers that the crystals will form asymmetrically due to entropically required chain folds and that the surface energy for the fold surface is much higher than that for the c-axis sides..

R
At the equilibrium melting point G = 0 = H - T S, so S = H/ T. At some temperature, T, below the equilibrium melting point, The volumetric change in free energy for crystallization fT = H - T S = H(1 - T/T) = H(T - T)/ T. The crystallite crystallized at "T" is in equilibrium with its melt and this equilibrium state is adjusted by adjusting the thickness of the crystallite using the surface energy, that is, GT = 4Rt side+ 2R2 - R2t fT = 0 at T. That is, At T the crystallite of thickness "t" is in equilibrium with its melt and this equilibrium is determined by the asymmetry of the crystallite, t/R. If fT = H(T - T)/ T. is use in this expression, 4t side+ 2R = R t H(T - T)/ T. Assuming that side <<< , and "t"<<<"R" then, t = 2 T./( H(T - T)) which is the Hoffman law. The deeper the quench, (T - T), the thinner the crystal and for a crystal crystallized at T, the crystallite is of infinite thickness. (Crystallization does not occur at T). Nature of the Chain Fold Surface: In addition to determination of T, the specific nature of the lamellar interface in terms of molecular conformation is of critical importance to the Hoffman analysis. There are several limiting examples, 1) Regular Adjacent Reentry , 2) Switchboard Model (Non-Adjacent
6

Reentry), 3) Irregular Adjacent Reentry (Thickness of interfacial layer is proportional to the temperature).

The synoptic or comprehensive model involves interconnection between neighboring lamellae through a combination of adjacent and Switchboard models.

The interzonal model involves non-adjacent reentry but considers a region at the interface where the chains are not randomly arranged, effectively creating a three phase system, crystalline, amorphous and interzonal. Several distinguishing features of the lamellar interfaces are characteristic of each of these models. Adjacent Uniform and Thin Fold Surface High Surface Energy Switchboard Random chains at interface, Broad interface, Low Surface Energy Irregular Adjacent Temperature Dependent interfacial thickness Intermediate Surface Energy Interzonal Extremely Broad and diffuse interfaces with non-random interfacial chains Synoptic Interfacial properties are variable depending on state of entanglement and speed of crystallization.
7

The Hoffman equation states that the lamellar thickness is proportional to the interfacial energy so we can say that Adjacent reentry favors thicker lamellae since adjacent reentry has the highest interfacial energy and the more random interfacial regions should display thinner lamellae. Colloidal Scale Structure in Semi-Crystalline Polymers: Lamellae crystallized in dilute solution by precipitation can form pyramid shaped crystallites which are essentially single lamellar crystals (figure 4.21 for example). Pyramids form due to chain tilt in the lamellae which leads to a strained crystal if growth proceeds in 2 dimensions only. In some cases these lamellae (which have an aspect ratio similar to a sheet of paper) can stack although this is usually a weak feature in solution crystallized polymers. Lamellae crystallized from a melt show a dramatically different colloidal morphology as shown in figure 4.30 pp. 182, 4.13 on pp. 157, 4.7 on pp. 149, 4.6 on pp. 148, 4.4 and 4.5 on pp. 147 and 4.2 on pp. 145. In these micrographs the lamellae tend to stack into fibrillar structures. The stacking period is usually extremely regular and this period is called the long period of the crystallites.
Long Period Amorphous C rystalline

The long period is so regular that diffraction occurs from regularly spaced lamellae at very small angles using x-rays. Small-angle x-ray scattering is a primary technique to describe the colloidal scale structure of such stacked lamellae. The lamellae are 2-d objects so a small angle pattern is multiplied by q2 to remove this dimensionality (Lorentzian correction) and the peak position in q is measured, q*. q= 4 / sin(/2), where is the scattering angle. Bragg's law can be used to determine the long period, L = 2/q*. Figure 4.8 on pp. 151 shows such Lorentzian corrected data. The peak occurs at about 0.2 degrees! In some cases the x-ray data has been Fourier transformed to obtain a correlation function for the lamellae which indicate an average lamellar profile as shown in figure 4.9 pp. 152. The degree of stacking of lamellae would appear to be a direct function of the density of crystallization, i.e. in lower crystallinity systems stacking is less prominent, and the extent of entanglement of the polymer chains in the melt. You can think of lamellar stacking as resulting from a reeling in of the lamellae as chains which bridge different lamellae further crystallize as well as a consequence of spatial constraints in densely crystallized systems. In melt crystallized systems, many lamellar stacks tend to nucleate from a single nucleation site and grow radially out until they impinge on other lamellar stacks growing from other nucleation sites. The lamellar stacks have a dominant direction of growth, that is, they are laterally constrained in extent, so that they form ribbon like fibers. The lateral constraint in melt crystallized polymers is primarily a consequence of exclusion of impurities from the growing crystallites.

Excluded

Impuriti es Fibrillar Growth Front

Nucleat ion Sit e

Excluded

Impuriti es

"Impurities" include a number of things such as dirt, dust, chain segments of improper tacticity, branched segments, end-groups and other chain features which can not crystallize at the temperature of crystallization. Some of these "Impurities" will crystallize at a lower temperature so it is possible to have secondary crystallization occur in the interfibrillar region. Despite the complexity of the "impurities" it can be postulated that the impurities display an average diffusion constant, D. The Fibrillar growth front displays a linear growth rate, G. Fick's first law states that the flux of a material, J, is equal to the negative of the diffusion constant times the concentration gradient c/x. If we make an association between the flux of impurities and the growth rate of the fibril then Fick's first law can be used to associate a size scale, x with the ratio of D/G. This approach can be used to define a parameter , which is known as the Keith and Padden -parameter, = D/G. This rule implies that faster growth rate will lead to narrower fibrils. Also, the inclusion of high molecular weight impurities, which have a high diffusion constant, D, leads to wider fibrils. There is extensive, albeit qualitative, data supporting the Keith and Padden del parameter approach to describe the coarseness of spherulitic growth in this respect. Branching of Fibrils: Dendrites versus Spherulites. Low molecular weight materials such as water can grow in dendrite crystalline habits which in some ways resemble polymer spherulites (collections of fibrillar crystallites which emerge from a nucleation site). One major qualitative difference is that dendritic crystalline habits are very loose structures while spherulitic structures, such as shown in Strobl, fill space in dense branching. At first this difference might seem to be qualitative.

120

In low-molecular weight materials such as snowflakes or ice crystallites branching always occurs along low index crystallographic planes (low Miller indices). In spherulitic growth there is no relationship between the crystallographic planes and the direction of branching. It has been proposed that this may be related to twinning phenomena or to epitaxial nucleation of a new lamellar crystallite on the surface of an existing lamellae. A definitive reason for non-crystallographic branching in polymer spherulites has not been determined but it remains a distinguishing feature between spherulites and dendrites. (Incidentally, the growth of dendrites can occur due to similar impurity transport issues as the growth of fibrillar habits in polymers. In some cases a similar mechanism has been proposed where rather than impurity diffusion, the asymmetric growth is caused by thermal transport as heat is built up following the arrows in the diagram on the previous page.) Non-crystallographic branching leads to the extremely dense fibrillar growth seen in figures 4.4 to 4.7 of Strobl. In the absence of non-crystallographic branching, many of the mechanical properties of semi-crystalline polymers would not be possible. As was mentioned above, non-crystallographic branching may be related to the high asymmetry and the associated high surface area of the chain fold surface which serves as a likely site for nucleation of new lamellae as will be discussed in detail below in the context of Hoffman/Lauritzen theory. The formation of polymer spherulites requires two essential features as detailed by Keith and Padden in 1964 from a wide range of micrographic studies: 1) Fibrillar growth habits. 2) Low angle, Non-crystallographic branching. Polymer Spherulites. Figure 4.2 pp. 145 shows a typical melt crystallize spherulitic structure which forms in most semicrystalline polymeric systems. The micrographs in figure 4.2 are taken between crossed polars and the characteristic Maltese Cross is observed and described on the following page. The Maltese cross is an indication of radial symmetry to the lamellae in the spherulite, supporting fibrillar growth, low angle branching and nucleation at the center of the spherulite. In some systems, especially blends of non-crystallizable and crystallizable polymers, extremely repetitive banding is observed in spherulites as a strong feature, figure
10

4.7 pp. 149. Banding is especially prominent in tactic/atactic blends of polyesters and it is in these systems in which it has been most studied. It has been proposed by Keith that banding is related to regular twisting of lamellar bundles in the spherulite (circa 1980). Keith has proposed that this twisting is induced by surface tension in the fold surface caused by chain tilt in the lamellae (circa 1989). Since most spherulites crystallize in an extremely dense manner it has been difficult to support Keith's hypothesis with experimental data. Regular banding has, apparently, no consequences for the mechanical properties of semi-crystalline polymers so has been essentially ignored in recent literature. XRD of Polymers: Four main features of XRD are of importance to Polymer Analysis: 1) 2) 3) 4) Indexing of Crystal Structures Microstructure Degree of Crystallinity Orientation

1) Indexing of Crystal Structures: Indexing of crystal structures is similar to the descriptions in Cullity and other metallurgical texts. The main difference is that polymer crystals can not be formed in perfect crystals, so single crystal or Laue patterns are not possible. Also, polymer crystals tend to be of low symmetry, orthorhombic or lower symmetry, due to the asymmetry in bonding of the crystalline lattice, i.e. the c-axis is bonded by covalent bonds and the a and b axis are bonded by van der Waals interactions or hydrogen bonds. Additionally, the unit cell form factor tends to be fairly complicated in polymer crystals. Several unit cells for polymers are shown below:

Nylon 66, from Alexander, "X-Ray Diffraction Methods in Polymer Science"

11

Polybutadiene (PBD), from Alexander, "X-Ray Diffraction Methods in Polymer Science"

Poly(ethylene adipate), a polyester, from Alexander, "X-Ray Diffraction Methods in Polymer Science" Lattice parameters in polymer crystals are strongly temperature dependent as shown in the following diagram:

12

From Balta-Calleja and Vonk, "X-ray Scattering of Synthetic Polymers" Polymer lattice parameters are also dependent on strain as shown in the following diagram:

From Balta-Calleja and Vonk, "X-ray Scattering of Synthetic Polymers" Notice that the c-axis (covalent main chain bonds) is much less dependent on thermal or mechanical strain. Line widths are broad for polymer diffraction and a substantial amorphous peak is usually present.

13

Isotactic Polystyrene, from Alexander, "X-Ray Diffraction Methods in Polymer Science" 2) Microstructure: Cullity deals with metallurgical crystals where crystallite sizes are typically larger than a micron. With a monochromatic incident beam the diffraction pattern from a single crystal is a sequence of spots where the Bragg condition is met for certain orientations of crystals (see "a" in figure below). As the crystallite size becomes smaller, more crystallites meet the Bragg condition and the radial orientation of these crystallites cover a broader spectrum of angles ("b" and "c" below), eventually forming Debye-Scherrer powder pattern rings ("c" below). If crystallite sizes approach 0.1 micron (1000), the Debye-Scherrer ring begins to broaden ("d" in figure below).

14

From Cullity, "Elements of X-Ray Diffraction Polymer crystals are on the order of 100 in thickness. Broadening of the diffraction lines due to small crystallite size becomes a dominant effect and the breadth of the diffraction lines can be used to measure the thickness of lamellar crystals using the Scherrer equation: t= 0.9 B cos( )

is the x-ray wavelength, B is the half width at half height for the diffraction peak in radians and is half of the diffraction angle. The Scherrer equation is derived in Cullity and other texts. Use of the Scherrer equation is a primary technique to determine lamellar thickness in polymer crystallites. This can be used in conjunction with the Hoffman-Lauritzen (Gibbs-Thompson) equation for studies of crystallization. In addition to Scherrer broadening diffraction lines can be broadened in polymers due to defects in the structure. This will not be covered in detail in this course but is described in Campbell and White and in Alexander's text. 3) Degree of Crystallinity: Polymers are never 100% crystalline since the stereochemistry is never perfect, chains contain defects such as branches, and crystallization is highly rate dependent in polymers due to the high viscosity and low transport rates in polymer melts. A primary use of XRD in polymers is
15

determination of the degree of crystallinity. The DOC is determined by integration of a 1-d XRD pattern such as that shown below for polyethylene.

From Balta-Calleja and Vonk, "X-ray Scattering of Synthetic Polymers" The determination of the degree of crystallinity implies use of a two-phase model, i.e. the sample is composed of crystals and amorphous and no regions of semi-crystalline organization. The alternative to the two-phase model is a paracrystalline model which was popular in the early days of polymer science. There are limits to the two-phase model, particularly for fairly disorganized polymer crystalline systems such as polyacrylonitrile (PAN). Most polymer systems are amenable to the two-phase model but you should keep in mind that the 2-phase model ignores interfacial zones where the density may differ from that of the amorphous. The integrated XRD intensity measures the volume fraction crystallinity, c. Other techniques such as density gradient columns (see Campbell and White or DSC) measure a mass fraction crystallinity c. The two fractions are related by the density ratios, where c is the crystalline density, is the bulk sample density and a is the amorphous density,

c =

c c

and

(1 c ) =

(1 c )a

16

If the density of the sample is known from a density gradient column, the weight fraction degree of crystallinity can be obtained using:

c =

c a c a

Determination of c from the XRD pattern under the 2-phase assumption involves separation the diffraction pattern into three parts, 1) Crystalline; 2) Amorphous and 3) Compton Background (Incoherent Scattering). The diffracted intensity if proportional to the amount of each of these contributions. Consider the 2-d diffraction pattern shown just above section 2) above. The 1-d diffraction pattern is a line cut through this pattern as shown below:

2-d Diffraction Patter n 1-d Slice of 2-d Pattern I q


The actual scattered intensity is related to a volume integral of the diffracted peak in the 2-d pattern, Vc Ic (q )dVq = q 2 Ic (q )dq
0 0

Then the volume fraction degree of crystallinity is given by the ratio of the integral of the crystalline diffraction intensity over the total coherent scattering, i.e. after subtracting the incoherent scattering:

c =

q I (q)dq
2 c

q [ I (q) I
2 0

Compton

(q)]dq

This procedure is called the Ruland method and is valid for: 1) Random crystallite orientation (Powder pattern) 2) 3-d crystalline ordering 3) Validity of integrals for finite angles of measure, i.e. there is a point in angle where the crystallinity is not significant to I(q)
17

4) Crystalline peaks can be separated from the amorphous halo The Ruland equation can be modified for crystalline defects as described in Campbell and White and Alexander. Usually the simple form given above is sufficient. The Ruland method is shown in the figures below where Iq2 is plotted as a function of q (or s).

From Alexander, "X-Ray Diffraction Methods in Polymer Science"

18

From Alexander, "X-Ray Diffraction Methods in Polymer Science" 4) Orientation: Orientation is covered in the later chapters of metallurgical diffraction texts such as Cullity (see figure below).

19

From Cullity, "Elements of X-Ray Diffraction Orientation is a more dominant effect in polymer samples especially processed plastics (see figure below). Orientation is a dominant feature in control of the mechanical and physical properties of polymers.

From Tadmor and Gogos, "Principles of Polymer Processing"

20

From Alexander, "X-Ray Diffraction Methods in Polymer Science"

21

From Alexander, "X-Ray Diffraction Methods in Polymer Science" There are a number of techniques for the quantification of orientation from diffraction data. Cullity describes the use of stereographic projections on a Wulff Net (shown below left). The Wulff net is useful if single crystals are studied and it is desired to determine the orientation with respect to the diffraction experiment such as in orientation of semi-conductor samples for cleavage. In most polymer applications it is desired to determine the distributions of orientation for a polycrystalline sample with respect to processing directions such as the direction of extrusion, (machine direction MD), the cross direction (CD) and the sample normal direction (ND). A more useful stereographic projection for these purposes is the polar net or pole figure (shown below right).

22

From Alexander, "X-Ray Diffraction Methods in Polymer Science" The pole figure is a slice across the equator of the sphere of projection with the MD usually defined at the top of the pole figure and either the CD or ND as the right side. Normals to planes are projected from the south pole to the point of intersection on the sphere of projection and where they cross the equatorial plane a point is plotted on the pole figure. A typical polar figure for a processed polymer is shown in the figure below for the (110) and (020) normals for the polyethylene orthorhombic crystalline structure. Notice that the plane normals appear as a topographical plot since there is a distribution in orientation. The (110) and (020) reflections are the two dominant peaks in the 1-d diffraction pattern for PE shown above (start of section 3).

23

From Balta-Calleja and Vonk, "X-ray Scattering of Synthetic Polymers" The following figure shows the type of qualitative analysis of orientation which can be performed using pole figures. Generally, pole figures are constructed by computer software which is part of a diffractometer capable of measurement of pole figures such as the Siemens D-500.

24

From Alexander, "X-Ray Diffraction Methods in Polymer Science" The pole figure can give a qualitative picture of orientation in a polymer sample. Quantitative measures of orientation can be obtained by considering a radial plot of diffraction data.

2-d Diffraction Patter n MD

TD I 0

36 0

The intensity for a given diffraction line (given 2) has two peaks as a function of radial angle, reflecting the two normals to the diffraction plane relative to the MD/TD plane. The Hermans orientation function can be calculated for a given plane from the dependence of the diffracted intensity:
25

f(110 ) =

1 3 cos2 1 2

where <cos2> is the average cosine squared weighted by the intensity as a function of the radial angle for the (110) plane. The Hermans orientation function has the behavior that f = 1 corresponds to perfect orientation in the = 0 direction, f = 0 for random orientation and f = -1/2 for perfect orientation normal to the = 0 direction. If the orientation function is calculate for orthogonal axis such as the a, b, and c unit cell directions for the PE unit cell then fa + fb + fc = 0. The orientation function for the unit cell vectors can be determined from geometry if the angular relationship between a plane normal and the unit cell direction is known. <cos2> is calculated by:

cos (110 ) =
2

I (, ) sin cos
(110 ) 0

I (, ) sin d
(110 ) 0

The figure below shows the behavior of the Hermans orientation function for the three unit cell directions in PE as a function of processing conditions in a fiber spinning process.

From Tadmor and Gogos, "Principles of Polymer Processing" The orientation function is directly related to polymer properties as shown in the example below.

26

From Tadmor and Gogos, "Principles of Polymer Processing"

27

Small Angle X-ray Scattering (SAXS) We have considered that Bragg's Law, d = /(2 sin ), supports a minimum size of measurement of /2 in a diffraction experiment (limiting sphere of inverse space) but does not predict a maximum size, i.e. the point (000) of inverse space reflects infinite size. In class we have discussed the use of diffraction to measure crystalline and amorphous structures on the atomic scale, but clearly, many morphologies are of importance that have characteristic sizes much larger than the atomic scale. In metals this was first noted by Guinier in his development of the theory of Guinier-Preston zones. Guinier was one of the fathers of an outgrowth of diffraction aimed at large-scale structures in the 1950's. Bragg's Law predicts that information pertaining to such nano- to colloidal-scale structures would be seen below 6 2 in the diffractometer trace. It is possible to design specialized instruments to measure down to less than 1/1000 of a degree for measurement of up to 1-micron scale structures using x-rays! The characteristics of materials at these larger size scales are fundamentally different than at atomic scales. Atomic scale structures are characterized by high degrees of order, i.e. crystals, and relatively simple and uniform building blocks, i.e. atoms. On the nano-scale, the building blocks of matter are rarely well organized and are composed of rather complex and non-uniform building blocks. The resulting features in x-ray scattering or diffraction are sharp diffraction peaks in the XRD range and comparatively nondescript diffuse patterns in the SAXS range. In XRD the scattered intensity depends on the Lorentz-Polarization factor which is essentially equal to 1 below 6 2. For disorganized systems the multiplicity factor is 1 and the structure factor, |F2|, generally does not reflect order so involves only a form factor for the nano-scale structures that give rise to scattering, i.e. regions of differing electron density. In XRD the atomic scattering factor, f2, was equal to the square of the number of electrons in an atom at low angles, n e2(1/q), where q is 4 sin( )/. Additionally, the intensity of scattering is known to be proportional to the number of scattering elements in the irradiated volume, Np(1/q). Then, in small-angle scattering we can consider a generalized rule that describes the behavior of scattered intensity as a function of Bragg size "d" or "r" that is observed at a given scattering angle 2, where r = 1/q. I(q) = Np(1/q) ne2(1/q) (1)

From this simplified rule of thumb we can derive most of the general rules of small-angle scattering in a less than rigorous manner. This approach, however, is extremely useful for a simple understanding of small-angle scattering. Scattering laws in the small angle regime describe two main features that are observed in a log Intensity versus log q plot. First, typical scattering patterns display power-law decays in intensity reflecting power-law scaling features of many materials. Secondly, power-law decays begin and end at exponential regimes that appear as knees in a log-log plot. These exponential knees reflect a preferred size as described by r = 1/q for the knee regime. All scattering patterns in the small-angle regime reflect a decay of intensity in q and this can be easily described by considering that at decreasing size scales the number of electrons in a particle is proportional to the decreasing volume, while the number of such particles increases

with 1/volume. Then the scattered intensity by equation (1) is proportional to the decay of the particle volume with size. This analysis implies that the definition of a particle, i.e. r, does not necessarily reflect a real domain, but reflects the size, r, of a scattering element that could be a component of a physical domain. Porod's Law: Next consider a sharp smooth surface of a particle such as a sphere. The surface can be decomposed into spherical scattering elements that bisect the particle/matrix interface. The number of such spheres is proportional to the surface area for the particles divided by the area per scattering element, r2 or 1/q2, while the number of electrons per particle is just proportional to r3 or 1/q 3. Using equation (1) with Np = Sq 2 and n e = 1/q 3, yields I(q) = S/q4, or Porod's Law for surface scattering. Porod's Law can be used to measure the surface area of domains in the nanoscale. The rigorously derived form of Porod's Law is, I(q) = Ie 2 2 S/q4 (2)

where is the electron density difference between particulate domains and the matrix material and Ie is a constant.

High Density Polyethylene showing XRD at high-q, SAXS at intermediate q and LS (light scattering) at low-q. Two Porod Regimes are observed.

Light Scattering and SAXS from non-woven mat of micron scale fiber glass. There are three other general categories of power-laws that are well defined in small-angle scattering, A) Surface-Fractal Laws, B) Diffuse interface Laws and C) Mass-Fractal or Dimensional Laws . Additionally there is the possiblity of D) polydispersity of particle size leading to power-law decays. A) Surface Fractal Laws. Systems that do not have smooth surfaces are often described by a scaling law where the surface area, S(r), is a function of the size of measurement. For example, the coastline of an island will increase if it is measured with smaller rulers since smaller rulers are capable of measuring the nooks and cranies of the coast line. For a smooth surface S(r) = r2, and for a rough surface S(r) = rds, where d s is the surface fractal dimension that varies from 2 to 3. Inserting such a scaling law into the discussion of Porod's Law above, yields I(q) proportional to qds-6. Surface fractals display power-law decays weaker than Porod's Law and are termed positive deviations from Porod's Law. B) Diffuse interfaces. Diffuse interfaces are observed when a concentration gradient is observed at an interface such as in mixing of two liquids or dissolution of a particle. A diffuse interface can be modeled as a power-law distribution of particles of differing electron density. Through this type analysis it is possible to describe the Schmidt exponent, , where I(q) goes as q-(4 + ), and describes the concentration decay of the diffuse interface.

C) Dimensional scattering laws. Particles can be described in terms of their dimension in the sense that a rod is 1-D, a disk 2-D and a sphere 3-D. Such a dimensional description implies that the mass of the object depends on the size of observation, "r", raised to the dimension. Because of this it is natural to expect power-law scattering to result form such low dimension objects. Consider that scattering from a rod is observed at "r" = 1/q between the rod length, L, and the rod diameter, D. Then the number of scattering elements in the rod of size "r" is equal to L/r or Lq. The number of electrons per scattering element is given by rD2 or D2/q. Using equation (1) we have I(q) = LD4/q, or a q-1 decay in intensity for the 1-D object. For a disk at size scales, r = 1/q, between the disk diameter, D, and thickness, t, the number of scattering domains is equal to the disk area, D2, divided by the scattering domain size, r2, or N p = D2/r2 = D2q2. The number of electrons per scattering domain is given by the volume of a domain, r2t= t/q2. Equation 1 yields I(q) = D2t2/q2, or I(q) proportional to 1/q2. Again the scattering is proportional to q-df where df is the dimension of the object. For a mass-fractal object such as a polymer coil the mass is given by the size raised to the dimension, i.e. for a polymer coil the end to end distance R is given by n1/2 l, so the mass, n is proportional to R2. In this sense a Gaussian polymer coil is a 2-D object. If such an object is broken into scattering elements of size r, the number of such elements is given by R2/r2 = R2q2 and the number of electrons in an element is ne = r2 = 1/q2. Then equation 1 yields I(q) = R2/q2, or I(q) is proportional to 1/qdf. D) Polydisperse Particles Many systems display dispersion in particle size in the small-angle regime. In some cases these dispersions are broad enough that they can be observed in terms of a power-law regime in scattering, i.e. when the dispersion in size occurs over a decade or more in size. Guinier's Law: We have so far discussed power-law decays in scattering that are mostly defined between two size limits, i.e. for a rod a power-law decay of q-1 is observed between the rod length and the rod diameter. Power-laws merge with a generic description of a discrete size at these limits, i.e. at q = 1/L or q = 1/D for a rod. For an isotropic system we can consider a particle in an average sense by allowing the particle structure to be averaged with respect to position and rotation. The scattering event involves interference from waves emanating from two points in the particle separated by a distance r = 1/q. Then the probability of constructive interference involves the probability that given a point lying in an average particle, one finds a second point also in the average particle at a distance r = 1/q. For an isotropic system one must consider first averaging any starting point in a particle and secondly averaging any direction for the vector r. This leads to a double summation that is identical to the determination of the moment of inertia for a particle. When the electron density is used as the weighting rather than the mass density, this moment of inertia is called the radius of gyration of the particle. For a system of disperse shaped
4

and sized particles the radius of gyration reflects a second moment of the distribution of the shape and size about the mean. The process of obtaining Rg involves two steps, first averaging all possible positions in the particle from which a vector "r" can start and be within the particle. Second, determining the probability that a randomly directed vector "r" from an arbitrary starting point in the particle will fall in the particle. The meaning of this probability p(r), in the vicinity of r = the particle size, can be graphically represented by a Gaussian probability cloud created by the summation of all possible positions of the particle where the center of the probability cloud is in the particle phase. At very low q this corresponds to the volume fraction particles squared. At sizes, r = 1/q, close to the average particle size or radius or gyration, this probability is reflected by a decaying exponential function. The decaying exponential function can be written in terms or r or in terms of q. (Fourier transform of a Gaussian distribution is a Gaussian distribution). Such an analysis leads to Guinier's Law, where the average size is reflected in the radius of gyration, Rg. Rg is the moment of inertia for the particle using the electron density rather than the mass as a weighting factor. I(q) = Np ne2 exp(-q2Rg2/3) Special Scattering Functions: Sphere: radius R I(q) = N n2{3(sinqR - qR cosqR)/(q3R3)}2 Rg = R/1.29 = R(3/5) Rod: Length 2H; Diameter 2R I(q) = N n2 exp(-q2R2/4) / (2qH) Rg overall = R2/2 + H2/3 Disk: diameter 2R; thickness 2H I(q) = N 2n2 exp(-q2H2/3) / (q2R2) Rg overall = R2/2 + H2/3 Gaussian Polymer Coil: n persistence units; l persistence length I(q) = I0 {2(Q - 1+exp(-Q)}/Q2 Q = q2Rg2
5

Rg overall = n1/2l/6 Unified Function: (one level) I(q) = G exp(q2Rg2/3) + B q*-P q* = q/(erf(qRg/6))3

This is a generic function where G, Rg, P and B are defined according to local functions. For example, for a sphere, G = N n2; R g = R/1.3; P = 4; B = 2 G S/V2 = 9 G/(2 R4). (Sphere S = 4R2; V = 4R3/3) Other examples can be found in: Small-Angle Scattering from Polymeric Mass Fractals of Arbitrary Mass-Fractal Dimension, Beaucage, G. , J. Appl. Crystallogr. (1996), 29, 134-146. Approximations leading to a unified exponential/power-law approach to small-angle scattering, Beaucage, G., J. Appl. Crystallogr. (1995), 28(6), 717-28. Correlation Function Analysis:

You might also like