Free Energy

The free energy (E) calculated from the Dubinin-Radushkevich isotherm model for Pb2+, Cu2+, and Cd2+ adsorption was between 12 and 19kJ/mol.

From: Sorbents Materials for Controlling Environmental Pollution, 2021



Related terms:




General applications

In Thermal Physics of the Atmosphere (Second Edition), 2021


3.1.3 Helmholtz free energy

The Helmholtz free energy, or simply free energy, is another thermodynamic potential. The specific free energy f (in the wider literature the letter a is also used) is defined as

(3.8)►�=�−��.

Specific free energy is an intensive variable; the free energy �=�−�� is an extensive variable. Its differential form again follows from the first law in differential form, analogous to the derivation of the differential of the enthalpy. We find

(3.9)d�=−�d�−�d�.

The free energy has as natural variables the temperature and the volume. Free energy plays a central role in statistical mechanics where it is natural to consider systems with a given temperature and volume, so that their free energy is fixed. A typical technique of statistical mechanics is to maximize the entropy S of a system (a measure of how probable a particular state is according to the Boltzmann definition of S) for a fixed total energy U. We then introduce a Lagrange multiplier β so that we maximize �−��. This can be interpreted as the (negative) free energy if β is interpreted as the inverse temperature. In this way, the microscopic world of statistical mechanics is linked to the macroscopic world of thermodynamics; see also Section 4.7. In atmospheric science free energy is perhaps used less often.

For the first derivatives of the free energy we find

(3.10)�=−(∂�∂�)�,�=−(∂�∂�)�.

The corresponding Maxwell relation is

(3.11)►(∂�∂�)�=(∂�∂�)�.



Energy processing by animals

David E. Reichle, in The Global Carbon Cycle and Climate Change, 2020


5.2 Free energy

Free energy or Gibbs free energy G, is the energy available in a system to do useful work and is different from the total energy change of a chemical reaction. Thus,

(5.3)Δtotalenergy=Δutilizableenergy+Δnon-utilizable energyΔH=ΔG+TΔSorΔG=ΔH−TΔS

where:

G is Gibbs free energy (kJ mol−1)

H is the heat of combustion, enthalpy (kJ mol−1)

T is temperature (°Kelvin), and

S is entropy (J°K−1)

To illustrate the concept of free energy, let us return to the familiar example of glucose oxidation. When 1 mol of glucose combines with 6 mol of carbon dioxide and 6 mol of water, the heat of combustion, ΔH, amounts to 673,000 calories:

(5.4)C6H12O6 + 6O2 = 6CO2 + 6H2O + 673,000 cal

Since there is no volume change, the total energy change, ΔE, is equal to the change in enthalpy, ΔH. The heat of formation, H, of glucose from the basic elements of CO2 and H2O can be calculated by subtracting the heat of combustion of glucose from the heats of formation of 6 mol of CO2 and 6 mol of H2O. The oxidation of 1 g-atomic-weight of solid carbon (graphite) to 1 mol of CO2 gas yields a heat of combustion, ΔH, of −94,240 cal. The heat of formation of 1 mol of water by the combustion of 1 mol of hydrogen gas with ½ mole of oxygen gas amounts to −68,310 cal. The heat of formation, H, of glucose is then calculated as:

6 mol CO2 (g) = 6 × −94,240 cal = −565,440 cal

6 mol H2O (L) = 6 × −68,310 cal = −409,860 cal

Heats of formation of products of glucose combustion: 975,300 cal

Heat combustion of 1 mol glucose (g) = −673,000 cal

Heat of formation of glucose from CO2 and H2O = −302,300 cal

The free energy change, ΔG, in the formation of glucose can be calculated from the previous equation:

(5.5)ΔG=ΔH−TΔS

Entropy values for the various atoms can be found in the Handbook of Physics and Chemistry. Essentially, the values given are relative heat capacities at very low temperature equivalent to absolute zero (−273°C):

(5.6)6×1.3+12×15.62+6×24.52=−342.4entropyunits(for C)(for H)(for O)

One gram of glucose represents −50.7 entropy units.

Entropyforglucose=elementaryentropyof =entropyinglucose
(5.7)ΔS298at 25°C=−342.1+50.7atmos in glucose=−291.4entropyunits

Therefore:

ΔH=−302,300calTΔS=298×−291.4=+86,500calΔG298=−215,800cal

This is the free energy change in glucose formation. The free energy involved in the oxidation of glucose is the difference between the free energy of formation of the combustion products and that of glucose, because by definition the free energy of O2 is zero.

G in 6 mols CO2 = 6 × (−94,100) =  = −564,600 cal

G in 6 mols H2O = 6 × (−56.560) =  = −339,360 cal

∑G in combustion products = −903,960 cal

G of glucose = −215,800 cal

ΔG in combustion of glucose = −688,166 cal

The change in free energy, ΔG, when glucose is oxidized is thus about 2% greater than the heat of combustion, ΔH, measured in a bomb calorimeter. This is because of the heat capacities of the reaction products and the fact that the reacting systems could also absorb heat from the environment. Although the absolute values of ΔH and ΔG are usually similar, conceptually they are quite different.

At this point you may be confused about why free energy values have been negative. Gibbs free energy is a derived quantity that blends together the two great driving forces in chemical and physical processes, namely enthalpy change and entropy change. Think of entropy as the measure of the random movement of molecules in the system. High entropy means a more random or chaotic state, such as a gas compared to a liquid. Processes in which entropy decreases tend not to occur in nature, unless there is a significant input of energy to cause them to take place.

Consider a reaction run at constant temperature, ΔG = ΔH−TΔS, where ΔH is the enthalpy change (the heat of the reaction) and ΔS is the change in entropy, If the free energy is negative, there is a change in enthalpy and entropy that favor the process and it occurs spontaneously. When Δ G is negative, the reaction is exergonic, and releases energy. This includes exothermic reactions in which the entropy increases, or exothermic reactions which have small decreases in entropy (as long as the temperature is relatively high), and endothermic reactions which are accompanied by large increases in entropy (like evaporation of water). When Δ G is negative, then the reaction is exergonic and releases energy.




Geological Sequestration of Carbon Dioxide

In Developments in Geochemistry, 2007


6.2.1 Nucleation and crystal growth

The precipitation of a solid phase comprises two distinct stages: nucleation, first, and crystal growth, afterwards. Nucleation is called homogeneous if the nuclei form in the bulk solution, whereas it is termed heterogeneous if they form on a solid surface. To describe the process, it is instructive to consider the variation of the free energy of precipitation, ΔGn, as a function of the radius of a hypothetical single crystal (Fig. 6.3).


Figure 6.3. Schematic diagram of the free energy of precipitation, ΔGn, vs. the radius, r, of a hypothetical single crystal. Note that ΔGn increases up to a maximum during nucleation and decreases afterwards during crystal growth. The effect of the degree of oversaturation, Ω, is also shown.

Reprinted from Stumm and Morgan (1996), modified, copyright (1996), with permission of Wiley.

The free energy of precipitation depends on the balance of two terms:

(i)

The interfacial free energy, which represents the energy associated with the formation of the interface between the growing crystal and the aqueous solution or, in other terms, the work required to generate this surface; this term is positive and depends on the square of the radius for a spherical crystal.

(ii)

The bulk free energy, which is negative and depends on the cube of the radius for a spherical crystal; besides, it is a function of the degree of oversaturation Ω (see Section 6.4 for the definition of Ω).

During nucleation (i.e. during the formation of the so-called crystal embryo), the interfacial free energy is greater than the bulk term and, consequently, the system experiences a net free energy increase. This increase occurs up to a maximum, corresponding to the balance between the interfacial and bulk terms. The growing solid phase, at this point, is called critical nucleus.

The further increase in the size of the critical nucleus is accompanied by a net decrease in free energy and the process becomes spontaneous for ΔGn < 0. This stage is known as crystal growth and the growing solid phase is referred to as a crystal. Of course, the transfer of material from the aqueous solution to the crystal goes on till the attainment of saturation (equilibrium).

The increase in Ω brings about both a decrease in the size of the critical nucleus and a decrease in the maximum of the ΔGn–radius curve (Fig. 6.3). Since, during nucleation, the rate is an exponential function of the free energy of the process, nucleation proceeds fast for high values of Ω and vice versa.

In the considered system, nucleation and crystal growth compete for the dissolved material. For high values of Ω, nucleation may proceed so fast that most dissolved material is used to constitute critical nuclei, whereas little is available for crystal growth. If so, a very fine-grained precipitate is formed. On the contrary, for low values of Ω, nucleation may proceed so slow that most dissolved material is consumed by crystal growth, involving a small number of critical nuclei. If so, a coarser mineral is produced.

In general, crystal growth comprises several mechanisms and its rate is limited by the slowest one. In particular, the rate of crystal growth may be governed by either

(i)

the transport of solute particles from the bulk aqueous solution to the crystal surface and, in this case, it is termed transport-controlled, or

(ii)

one of the numerous reactions taking place at the surface of the growing crystal and, if so, it is called surface-reaction controlled or

(iii)

by a combination of these two different mechanisms.

The change in the concentration of the relevant solute(s) moving away from the crystal surface towards the bulk aqueous solution is different, depending on the mechanism controlling crystal growth, as shown schematically in Fig. 6.4.


Figure 6.4. Plots (a) and (b) show schematically the change in the concentration of a relevant solute as a function of the distance from the crystal surface for (a) transport-controlled dissolution and precipitation and (b) surface-reaction-controlled processes. Plots (c) and (d) represent the change in concentration as a function of time in a generic batch experiment for (a) transport-controlled and (b) surface-reaction-controlled dissolution.

Reprinted from Stumm and Morgan (1996), modified, copyright (1996), with permission from Wiley.

In transport-controlled crystal growth (Fig. 6.4a), addition of solute particles to the solid surface (i.e. the surface reaction under way) is so quick that solute transport in the aqueous phase cannot keep up with it. Therefore, solute concentration in the aqueous solution near the solid surface decreases sharply and gets close to the saturation (equilibrium) value. Solute transport in the aqueous solution takes place through either advection, which is the fastest way, or diffusion, which is slowest way. Transport-controlled crystal growth depends on hydrodynamic conditions and stirring accelerates it.

In surface-reaction controlled crystal growth (Fig. 6.4b), addition of solute particles to the solid surface is so slow that transport processes, even diffusion, are able to supply new solute particles near the growing crystal surface. In this case, there is little change in solute concentration between the aqueous layer close to the solid surface and the bulk solution, and crystal growth is virtually independent of hydrodynamic conditions.

Intermediate situations are possible depending on the relative speed of solute addition to the crystal surface and solute transport in the aqueous phase. Both mechanisms govern the rate of crystal growth.

To identify the mechanism controlling the rate of crystal growth, experimentally determined rates are compared with those computed for the slowest type of aqueous transport, i.e. molecular (ionic) diffusion. Measured rates faster than diffusion-controlled rates are evidently explained by advective transport, whereas slower measured rates suggest that crystal growth is governed by reactions occurring at the solid surface.

Following Nielsen (1964), the diffusion-controlled rate is computed by means of the following equation:

(6-51)�����=�⋅��⋅(��−��)��

where rc is the mean radius of the crystals; ν the molar volume of the precipitating substance; DS the diffusion coefficient of solute particles in the aqueous solution; CB and CS are the concentrations of the solute in the bulk aqueous solution and close to the crystal surface, respectively; and t time. Assuming CB to be constant, integration of equation (6-51) gives

(6-52)��=[��,�=02+2⋅�⋅��⋅(��−��)⋅�]1/2

where rc, t=0 is the average radius of the crystals at time “zero, ” i.e. at the beginning of crystal growth, i.e. at the end of the nucleation step. equations (6-51) and (6-52) apply to equidimensional crystals of regular shape (e.g. spheres, cubes, etc.) separated by at least five diameters.

Alternatively, crystal growth is experimentally carried out at different temperatures and the temperature dependence of the rate is established to infer the controlling mechanism (see Section 6.1.3). Besides, dependence on hydrodynamic conditions suggests transport control and vice versa, as already mentioned.




Ring-Opening Polymerization and Special Polymerization Processes

W.J. Feast, in Polymer Science: A Comprehensive Reference, 2012


4.26.2.4.1 General considerations

The free energy for ring opening of cyclopentene is considerably less favorable than that of cyclobutene and, depending on the position and kind of substitution on the ring, ΔG can be just positive or just negative.15,16 Nevertheless, cyclopentene and many of its derivatives do undergo ROMP readily, polymer formation being favored by lower temperatures and higher monomer concentrations. At one point several years ago, the homopolymer from cyclopentene, polypentenamer, was under serious consideration as a substitute for commercial diene elastomer manufacture; indeed, it was reported that prototype vehicle tires were made and road tested before changes in economics and tire technology made the venture nonviable.

The mnemonic in Figure 4 indicates that the cyclopentadienyl radical is not a stable entity and will display a strong tendency to pick up an electron and become an aromatic cyclopentadienyl anion, as is indeed the case. So the direct route from cyclopentadienyl radical portrayed in Figure 3 is out of the question and routes to conjugated polymers involving ROMP of cyclopentene derivatives will inevitably involve other steps in the overall scheme. The ring-opening polymerizability of five-membered rings is enhanced by strain and this can be induced by making them part of a polycyclic structure, vide infra.






Basic Aspects of Radiation Effects in Solids/Basic Aspects of Multi-Scale Modeling

W.G. Wolfer, in Comprehensive Nuclear Materials, 2012


1.01.7.4.4 Chemical potential of vacancies at cavities

The free energy of a void or bubble, according to eqn [127], depends now on three surface parameters instead of just one as in eqn [105]: it depends on the surface energy γ0 of a planar surface, on the residual surface strain ε* for such a planar surface, and on the biaxial surface stretch modulus 2(μS + λS). As mentioned above, the latter has the dimension of N m−1, and we may then relate it to the corresponding bulk modulus 2μM/(1–2νM) by multiplying the latter with a surface layer thickness parameter h. The surface energy γ0 has been determined both experimentally and from ab initio calculations and can be considered as known. The surface layer has been determined by Hamilton and Wolfer10 from atomistic simulations on Cu thin films to be one monolayer thick; hence d = b. A value for the residual surface strain parameter ε* has been chosen in Section 1.01.3.1 such that it reproduces the relaxation volume of a vacancy according to eqn [11].

What if one selects the same value for voids containing n vacancies? The relative relaxation volume, that is, the ratio ��VR/(�Ω)=3ɛ(�(�)), can now be computed with eqn [123] and the results are shown in Figure 23 by the solid curve. As it must, for n = 1 it reproduces the vacancy relaxation volume of −0.25Ω. In addition, it also agrees with the overall trend of the atomistic results of Shimomura.48 Of course, the atomistic results for small vacancy cluster vary in a discontinuous manner with the cluster size. The surface stress model gives not only a reasonable approximation to these atomistic results, but also a valid extrapolation to relaxation volumes of large voids.



The chemical potential of vacancies for voids can now be computed with eqn [127] as FC(R(n + 1)) − FC(R(n)). Figure 24 shows the results for Ni as the solid curve. The vacancy chemical potentials for voids are significantly lower than the capillary approximation predicts with a fixed surface energy (dashed curve). The chemical potentials from atomistic simulations of voids in Ni have been obtained by Adams and Wolfer49 using the Ni-EAM potential of Foiles et al.50 These results converge to those predicted with the surface stress mode




New Techniques for Optimization of Particulate Cleaning

Per M. Claesson, ... Andrew Fogden, in Handbook for Cleaning/Decontamination of Surfaces, 2007


2.2. Surface Force Techniques and Detachment Force Measurements

The free energy change per unit area (ΔG) accompanying the separation of a surface from another can be directly measured with a range of different techniques, such as AFM and the surface force apparatus, SFA. The principles of such force measuring techniques will not be discussed here, but the interested reader is referred to the review by Claesson et al. and references therein [4]. By means of the Derjaguin approximation, the force (F) measured between a flat surface and a sphere with radius R is equal to the force between two crossed cylindrical surface (F) with a geometrie mean radius of R, and related to the free energy of interaction per unit area between flat surfaces at the same separation (D) as [5,6]:

(4)�(�)2��=�(�)flat

This relation is valid provided R ≫ D and provided surface deformation effects can be ignored. There are numerous reports in the literature describing forces acting between solid surfaces in Surfactant solutions, see e.g. [7] and references therein. Both long-range forces and contact forces are reported in the literature. For instance, the adhesion force in air is determined by bringing the surfaces into contact in air, and then separating them. Likewise, the adhesion force in water is determined by bringing the surfaces together in water and then separating them from contact. In a similar manner, the adhesion force between adsorbed Surfactant layers is determined in aqueous Surfactant solutions. The information obtained from such measurements is of great value. However, in most cases these measurements do not mimic the detachment process in a typical cleaning process sufficiently.

In a typical cleaning situation, the particle attaches to the surface in air and is removed in a liquid cleaning formulation. The nature of contact in the dry state is different from the nature of contact in the wet state, and a manifestation of this is the well-known observation from industrial tests that it is much more difficult to remove particles that have dried onto a surface than it is to remove particles that attach in the wet state and never is allowed to dry onto the surface. In contrast to this typical situation in a cleaning process, the measurements of adhesion forces in air, as carried out with e.g. the SFA, mimics the situation when a particle attaches and is removed in air. Similarly, the adhesion force measured in water by bringing the surfaces together from a large separation and then separating them relates to the process, of adsorbing and removing a partiele in water. In order to relate the surface force measurements to a typical cleaning process, a different measuring strategy has to be adopted. The surfaces are brought together in air, and then a droplet of an aqueous solution is placed around the contact region. Finally, the surfaces are separated within this droplet. The force needed to separate the surfaces under these conditions will be referred to as the detachment force in order to distinguish it from the adhesion forces (also called pull-off forces in the literature) normally measured with surface force techniques. The process described above can easily be adopted using the SFA, but the measurements are rather time consuming since only one measurement of the detachment force can be obtained in each experiment. On the subsequent approach the surfaces will be wetted, and on subsequent removal, the adhesion force in the liquid will be measured instead of the detachment force.

One may of course ask if the values of the adhesion force and the detachment force are significantly different, and if a distinction between these two quantifies is meaningful. The answer is yes. For instance, in the case of two muscovite mica surfaces the adhesion force in dry air is about 1100 mN/m [8], and the adhesion force in water is about 30–50 mN/m [9]. In contrast, the detachment force in water is about 300 mN/m, i.e. more than a factor of three lower than the adhesion force in air, and about a factor of 10 larger than the adhesion force in water! The detachment force is lower than the adhesion force in air due to water adsorption that lowers the interfacial energy. The reason that it is larger than the adhesion force in water is that when the surfaces are brought into contact in water some water will remain between the surfaces also when they are “in contact”, i.e. the contact situation is different in the two cases [10]. This is illustrated by the schematic force curve in Figure 8.2. The detachment force is the magnitude of the force at point D, whereas during normal force measuring procedures the force is measured from large distances,over the force barrier (A), into an adhesive minimum at B (the magnitude of which is the adhesion force), and further in another very large force barrier is encountered (C). In most experiments, this force barrier is not overcome and the “contact” achieved is not absolutely dry (except when hydrophobic surfaces are used).



No systematic study of detachment forces in presence of Surfactants and polymers has yet been published. However, the relation between detachment forces and cleaning has been investigated during the past years within the competence center “Surfactants based on natural products, SNAP”. One generai conclusion that has emerged from these studies is that the detachment force is smaller than the adhesion force in air and larger than the adhesion force in water in all cases investigated so far. We note that the adhesion force between Surfactant layers as measured with surface force techniques is typically between zero and a few milli newton per meter, whereas the detachment force between one hydrophilic mica surface and one hydrophobized surface in different Surfactant Systems varies significantly more, from close to zero in a few exceptional cases upto typical values in the order of 10 mN/m. We have also found a satisfac-tory correlation between a low detachment force and good performance in industrial tests.

Some selected data where the detachment force is expressed in fractions of the adhesion force in air are shown in Figure 8.3. We note that the anionic Surfactants linear alkylbenzene sulfonate (LAS) and sodium dodecyl sulfate (SDS) show very similar effect on the detachment force. The anionic Surfactants alone are slightly more efficient than the non-ionic Surfactants alone in their lowering of the detachment force with one exception, the branched nonionic glucoside C2C6Glu that gives rise to a lower detachment force. The data in Figure 8.3 also show how addition of polymers can be beneficiai in lowering the detachment force. One polymer studied, called M4, is weakly cationic and carries graftedpoly(oxyethylene) chains. Alone, it is not able to reduce the detachment force significantly. However, when combined with an anionic Surfactant the mixture that contains only 20 ppm of polymer, reduces the detachment force significantly more than the polymer alone and the Surfactant alone.


Figure 8.3. Detachment force divided by the adhesion force in air. The detachment force was measured at a surfactant concentration above the cmc using one hydrophilic mica surface and one hydrophobized mica surface. The surfactants were: C12E5, penta(oxyethylene) dodecyl ether; Cl0Glu, n-decyl β–d glucopyranoside; C2C6Glu, 2–ethylhexyl α–glucoside; Sterol ethoxylate, a sterol backbone connected with a 25 unit long poly(oxyethylene) chain; LAS, linear alkyl sulphate; SDS, sodium dodecyl sulphate, M4 a weakly cationic polymer with grafted EO–chains. The concentration of the polymer was 20 ppm


The reduction in the detachment force in presence of a Surfactant can be understood by considering that adsorption reduces the interfacial energy as described by the Gibb's equation. If we assume that the Surfactants do not adsorb in the gap between the surfaces, then the particle-surface interfacial energy is not affected. (This is confirmed by the SFA measurements which show that the contact position is not affected by introduction of the droplet around the rim of the contact zone). On the other hand, the Surfactants are likely to adsorb on the surfaces of the partiele and the substrate, reducing their interfacial energy with water. The relation between the detachment force in the Surfactant solution and in water is then given by;

(5)(�2��)sol=(�2��)w−∫�(�=0)�(�=�')�s��−∫�(�=0)�(�=�')�p��

where the subscripts “sol” stands for Surfactant solution, “w” for water, “s” for substrate surface and “p” for partiele surface. Thus, a Surfactant that is efficient in lowering the detachment force should adsorb strongly to both the substrate surface to be cleaned and the particulate soil surface. In particular, the larger the integral of the adsorption isotherms the larger the reduction in detachment force.

Surface force techniques are also useful for unraveling the events occurring during removal of proteins [11], polymers [12] and polyelectrolytes [13] from surfaces by addition of Surfactants. In some cases, large swelling of the layer occurs prior to desorption [14], in other cases a graduai thinning of the layer is observed [15], and in still other cases complex polymer–surfactant structures are formed at the surface [16]. It can also be the case that the Surfactant mainly adsorb on top of the polyelectrolyte layer with minimal reduction in the adsorbed mass of polymer [[17],[18].




Polymer Properties

Richard A. Brown, ... Xue Feng Yuan, in Comprehensive Polymer Science and Supplements, 1989


6.3.4 Thermodynamic Studies

Standard free energies of micellization ΔG and their enthalpy ΔH and entropy –TΔS components have been determined by Price et al. for a number of block copolymers in organic solvents using light scattering, membrane osmometry and calorimetry.153–158

One of the systems studied153 was a polystyrene-block-poly(ethylene/propylene) (37 300:59 700 Mncopolymer in decaneElectron microscopy studies showed that the micelles formed by the block copolymer were spherical in shape and had a narrow size distribution. Since decane is a selectively bad solvent for polystyrene, the latter component formed the cores of the micelles. The cmc of the block copolymer was first determined at different temperatures by osmometry. Figure 13 shows a plot of π/cRT against c (where c is the concentration of the solution) for T = 97.1 °C. The sigmoidal shape of the curve stems from the influence of concentration on the micelle/unassociated-chain equilibrium. When the concentration of the solution is very low most of the chains are unassociated; extrapolation of the curve to infinite dilution gives �n−1 of the unassociated chains. On increasing the concentration of the solution the osmotic pressure decreases rapidly over a narrow concentration range as expected for closed association. The arrow indicates the cmc. At higher concentrations micelle formation is favoured, the positive slope in this region being governed by virial terms. Similar shaped curves were obtained for other temperatures.



The slope of the linear plot of ln(cmc) against T−1 (see Figure 14) gave a value for ΔH of −130 kJ mol−1. The values of ΔG and TΔS for T = 86 °C were −30 and 100 kJ mol−1 respectively; the standard states are ideally dilute with c = 1 mol dm−3.



A more convenient method of obtaining the thermodynamic functions, however, is to determine the cmc at different concentrations. A plot of light-scattering intensity against concentration is shown in Figure 15 for a solution of concentration c = 3.8 × 10−5 g cm−3 and a scattering angle of 60°. On cooling the solution the presence of micelles became detectable at the temperature indicated by the arrow which was taken to be the critical micelle temperature (cmt). On further cooling the weight fraction of micelles increases rapidly leading to a rapid increase in scattering intensity at lower temperatures till the micellar state predominates. The slope of the linear plot of ln c against (cmt)−1 shown in Figure 16, which is equivalent to the more traditional plot of ln(cmc) against T−1, gave a value of ΔH = −141 kJ mol−1 which is in fair agreement with the result obtained by osmometry considering the difficulties in locating the cmc by the osmometric method. Direct calorimetric measurements gave a value of 138 kJ mol−1 for ΔH.


Figure 15. A plot of light-scattering intensity against temperature for a polystyrene-block-poly(ethylene/propylene) (37 000:59 700 Mn) copolymer in decane. The concentration of the solution was 3.85 × 10−5 gcm−3 and the angle of scatter 60°. The arrow indicates the cmt (reproduced from ref. 153)




Results obtained for a range of polymers are given in Table 3.154, 155, 159 The first two sets of results were obtained using light-scattering to determine the cmt. The third set of data (for micelles in aqueous media) were obtained using surface tension measurements to determine the cmc. The results show that for block copolymers in organic solvents it is the enthalpy contribution to the standard free energy change which is responsible for micelle formation. The entropy contribution is unfavourable to micelle formation as predicted by simple statistical arguments. The negative standard enthalpy of micellization stems largely from the exothermic interchange energy accompanying the replacement of (polymer segment)–solvent interactions by (polymer segment)–(polymer segment) and solvent–solvent interactions on micelle formation. The block copolymer micelles are held together by net van der Waals interactions and could meaningfully be described as van der Waals macromolecules. The combined effect per copolymer chain is an attractive interaction similar in magnitude to that posed by a covalent chemical bond.

Table 3. Thermodynamics of Micellizationa

Polymer Mb Mc ΔG (kJ mol−1) ΔH (kJ mol−1) TΔS (kJ mol−1)
Polystyrene-block-polyisoprene copolymerd 24 600e 7000e −21.4 −40.7 19.3
46 600e 8650e −23.6 −72.6 49.0
26 800e 13 000e −31.8 −86.4 54.6
54 228e 12700e −30.5 −115.3 84.8
Polystyrene-block-poly(ethylene/propylene) copolymerf 146 000g 49 800g −43.0 −181 138
97 000g 37 300g −41.5 −141 100
110 000g 31 500g −41.7 −103 61
Poly(oxyethylene) n-alkyl ethersh 1370g 150g,i −24.6 10.9 −35.5
1430g 210g,i −26.8 16.9 −43.7
1470g 250g,i −28.8 24.5 −53.3
1510g 290g,i −30.2 32.0 −62.2
1570g 350g,i −31.4 39.4 −70.8
a
Standard states are ideally dilute with c = 1 mol dm−3.
b
Overall average molecular weight of copolymer.
c
Average molecular weight of polystyrene block.
d
In hexadecane at T = 40 °C.154
e
M = Mw.
f
In decane at T = 80 °C.155
g
M = Mw.
h
In water at T = 35 °C.159
i
Average molecular weight of n-alkyl block.

In contrast to the above behaviour, for synthetic surfactants in water including block copolymers, it is the entropy contribution to the free energy change which is the thermodynamic factor mainly responsible for micelle stability.159, 160 Results for the thermodynamics of micellization of poly(oxyethylene) n-alkyl ethers (structural formula: MeO(CH2CH2O)27(CH2)nH, where n = 12, 14, 16, 18, 21) in water are given in Table 3. Whilst a number of factors govern the overall magnitude of the entropy contribution, the fact that it is favourable to micelle formation arises largely from the structural changes161 which occur in the water matrix when the hydrocarbon chains are withdrawn to form the micellar cores.















LINK: https://www.sciencedirect.com/topics/earth-and-planetary-sciences/f...

Views: 65

Have questions?

Need help? Visit our Support Group for help from our friendly Admins and members!

Have you?

Become a Member
Invited Your Friends
Made new Friends
Read/ Written a Blog
Joined/ Created a Group
Read/ Posted a Discussion
Checked out the Chat
Looked at/Posted Videos
Made a donation this month
Followed us on Twitter
Followed us on Facebook

Donations & Sponsorship

~~~~~~~~~~~
Please consider a donation to help with our continued growth and site costs

Connect

Visit The Temple
on Facebook:

....

Blog Posts

coat of arms of McIntyre clan.

Posted by Ghillie Dhu on March 2, 2024 at 5:06am 4 Comments

Are Ouija Boards Evil?

Posted by Bill Walker on February 1, 2024 at 8:15pm 1 Comment

TO RISE AND FIGHT AGAIN

Posted by Kitt on December 14, 2023 at 8:55pm 2 Comments

Osiris the Warlock: The Cursed

Posted by 06iiris on October 13, 2023 at 8:30am 0 Comments

Osiris the Warlock: Horus Lives!

Posted by 06iiris on October 13, 2023 at 7:00am 0 Comments

Osiris the Warlock: Nanna-Sin

Posted by 06iiris on October 12, 2023 at 3:30pm 0 Comments

Osiris the Warlock: Daemon

Posted by 06iiris on October 12, 2023 at 6:00am 0 Comments

Osiris the Warlock

Posted by 06iiris on October 12, 2023 at 5:30am 0 Comments

The (3l)ack & Red Dragon

Posted by 06iiris on October 11, 2023 at 1:00pm 0 Comments

The Sin (3l)ood Omen (II)

Posted by 06iiris on October 11, 2023 at 8:30am 0 Comments

The Sin (3l)ood Omen

Posted by 06iiris on October 9, 2023 at 7:30am 0 Comments

The 3nki (Doc)trine: Ptah-Øsiris

Posted by 06iiris on October 5, 2023 at 1:00pm 0 Comments

NOVEMBER AWARENESS

© 2024   Created by Bryan   Powered by

Badges  |  Report an Issue  |  Terms of Service