Chapter 6

Thermodynamics

6.1 Atomic Model

6.1.1 Protons, Neutrons, and Electrons

Atom. The name atom is derived from the Greek word ατομοϛ (atomos) that literally translates to “indivisible” or “not to be cut”. The concept of matter being ultimately composed of elements that may not be further divided was first brought forth by Dalton.1

Proton and Neutron. Atoms are composed of a very dense nucleus that consists of neutrally charged neutrons and positively charged protons. The positive charge of a proton amounts to one elementary charge that is equal to 1.602 176 565 × 10−19 C. The mass of protons is 1.672 621 777 × 10−27 kg. The mass of neutrons is 1.674 927 351 × 10−27 kg and they are usually considered to have approximately the same mass. The nucleus is held together by electromagnetic forces.

Electron. The electrons are arranged in shells around the nucleus. Electrons are three orders of magnitude lighter than both protons and neutrons and carry one negative charge. Its weight is 9.109 382 91 × 10−31 kg. An atom is classified by the number of protons, in electrically neutral states. If the number of electrons is equal to the number of protons (which is the default state), the atom is referred to as being electrically neutral or uncharged.

Ion. If the number of electrons and protons is not equal, the atom becomes charged. It is then referred to as an ion. As an example, Na+ indicates that the sodium atom has acquired one positive elementary charge or, to be correct, it lost one electron. The atom is then referred to as a cation. If the atom gained one elementary charge, it is referred to as an anion. An example is the chlorine anion Cl. Higher charges may, of course, also occur, e.g., in copper sulfate CuSO4 where the copper ion is double-positively charged Cu2+ and the sulfate is double-negatively charged SO42si8_e. Uncharged atoms can be transformed into ions by removing or adding electrons. Removal of electrons requires a certain amount of energy that is referred to as ionization energy. Thus the process of removing an electron from an atom is referred to as ionization, whereas the inverse process is referred to as recombination.

6.1.2 Periodic Table

Element. The number of protons in the nucleus defines the chemical element. Per definition, an element is a substance that cannot be decomposed any further. The number of protons is referred to as the atomic number that identifies the respective chemical element. The elements are arranged in the periodic table of the elements (see Fig. 6.1) in order of increasing atomic number. For each element, the number of protons and the number of neutrons are the same, with the exception of hydrogen, which has one proton, but no neutrons. The first composition of all known chemical elements was provided by the French chemist Antoine de Lavoisier [2]. It contained 33 substances. Today, we know 118 elements, 88 of which occur naturally.

f06-01-9781455731411
Fig. 6.1 Periodic Table of the elements.

Nuclide. If an atom is characterized by its number of protons and neutrons, it is referred to as a nuclide. Some elements have more than one nuclide, i.e., they occur with different neutron count. Such nuclides are referred to as isotopes of the pure element, i.e., as versions of the atom with varying number of neutrons, but the same number of protons. Isotopes are denoted with two preceding numeric values. The upper integer denotes the sum of neutrons and protons in the nucleus, the lower integer, the number of protons: 11Hsi9_e is protium (the naturally occurring form of hydrogen with one proton only), 12Hsi10_e is deuterium (commonly used hydrogen substitute in chemical spectroscopy with one proton and one neutron), 13Hsi11_e is tritium (the radioactive isotope of hydrogen with one proton and two neutrons).

Many elements occur as mixtures of their isotopes naturally. Examples include carbon, which occurs besides the normal 612Csi12_e (99 %) also as 613Csi13_e (1 %), that is important, e.g., in nuclear magnetic resonance (NMR) spectroscopy. Only a few elements, e.g., fluorine or sodium occur as only a single isotope. These elements are referred to as mononuclidic element.

The periodic table is arranged in periods and groups. Elements that belong to the same group in the periodic table have similar physical and chemical properties, as they share the same number of so-called valence electrons, i.e., electrons in the outmost shell. Elements from the same period vary with respect to their physical and chemical behavior but within periods, certain trends in their behavior can be observed.

Period. Within the same period (going from left to right):

 atom radii decrease, due to the fact that a higher number of protons in the atom nucleus exerts higher attractive forces on the electrons, thus pulling them closer to the nucleus

 ionization energy increases due to the increasing interaction forces between the nucleus and the electrons

 electronegativity (which is the ability to attract electrons) increases due to the fact that an increasingly positively charged nucleus exerts higher attraction forces

Group. Within the same group (going from top to bottom):

 atom radii increase due to the fact that the atom has more electrons

 ionization energy decreases as the attractive forces from the nucleus decrease, thus making it easier to remove electrons

 electronegativity decreases as the positively charged nucleus is further apart from electrons in close proximity due to the fact that the atom radius is bigger

By convention, the groups in the periodic tables are assigned certain names:

 group 1 - “alkali metals”

 group 2 - “alkaline earth metals”

 group 3 to 12 - “transition metals”

 group 13 - “boron group elements”

 group 14 - “carbon group elements”

 group 15 - “pnictogens”

 group 16 - “chalcogens”

 group 17 - “halogens”

 group 18 - “noble gases”

Please note, that sometimes periodic tables may additionally subdivide atoms into categories, which comprise atoms from different periods and groups with similar physical or chemical properties. These categories include, e.g., post-transition metals or metalloids. In the periodic table in Fig. 6.1 these categories are color-coded. As this categorization is not standardized, periodic tables from different sources may vary.

6.2 Weights and Concentrations

Mole and the Avogadro Constant. The universal weight reference in chemistry is the relative atomic weight u. It is equivalent to 112si14_e of the mass of the 612Csi12_e isotope. This is a very small number, 1.660 538 921 × 10−27 kg to be precise. In order to avoid calculating with this number, the symbol u is used instead. As chemistry usually handles quantities that are physically practical (after all, handling individual molecules is not an easy task) we assume a “useful” number of molecules. In principle, any number would do, but given that u already introduces a number with a negative exponential, why not take a number that gets rid of the exponential all together? So we define this number to be

NA=1u6.022×10231/mol

si16_e  (Eq. 6.1)

The number we have introduced here is referred to as the Avogadro constant. So if we take this number of molecules and calculate the overall weight of this mass, the result would amount to a round number. The number of molecules defined by the Avogadro1 constant is referred to as mole: 1 mol. Using the derivation of u, we can now see that 1 mol of the 612Csi12_e isotope has a total weight of 12 g, a much more practical number to work with. Using the Avogadro constant, the weight of a molecule is then expressed as the molecular weight in g mol−1. Sometimes, especially in biology and biochemistry, molecular weights are expressed in Dalton (Da). 1 Da is equivalent to 1 g mol−1 and the units can be used interchangeably in principle. However, especially when units are to be considered, g mol−1 uses the units conformal to SI and is therefore preferred. The Avogadro constant is often confused with the Loschmidt number NL that describes (for an ideal gas) the number of atoms or molecules contained in a given volume (see also section 6.4.1 for details on the Loschmidt number).

Molecular Weights. Now, calculating the molecular weight of a molecule is simply a question of adding up the values given for the relative atomic weights of the individual atoms that a molecule is composed of. These values are given, e.g., in the periodic table of elements (see Fig. 6.1) in kg, i.e., in relative atomic mass. Now how does this help us? Remember, the relative atomic mass is given as 112si14_e of the mass of the 612Csi12_e isotope in atomic units u. Now 1 mol of carbon weighs exactly 12 g; therefore, the molecular weight of carbon is equivalent to the relative atomic mass. Therefore, the molecular weight can be simply read from the periodic table.

As an example, the molecular mass of carbon dioxide can be calculated as

MCO2=MC+2MO=12gmol1+216gmol1=44gmol1

si20_e

As a second example, let’s consider the molecular mass of the salt potassium chloride

MKCl2=MK+2MCl=39gmol1+3545gmol1=74.45gmol1

si21_e

Using the molecular weight, we can formulate the dependence of the amount of substance (measured in number of molecules, i.e., in mol) and the mass of substance we take (given in g) as

n=mM

si22_e  (Eq. 6.2)

This allows us to convert from a given mass m of a substance to the number of molecules that are contained in this quantity using the molecular weight M. In general, chemistry works with substance amounts n rather then with mass m. This is due to the fact that chemical reactions involve individual molecules and if a chemical reaction requires two different molecules, it makes sense to supply a quantity that will ensure that there is a sufficient and balanced number of both molecules. As an example, in order to create carbon dioxide, one needs two molar parts of oxygen and one molar part of carbon. Therefore we need to supply 12 g carbon and 36 g oxygen in order to create 1 mol carbon dioxide. This correct balancing of substance amount is referred to as stoichiometry.

Uneven Atom Mass. As stated in section 6.1.2 some elements have more than one isotope in nature. This is common, e.g., for the noble gases. As an example, for neon the given atomic mass is 20.18 g mol−1. Neon has several isotopes: Ne1020(91%)si266_e, Ne1021(0.2%)si267_e, and Ne1022(8.8%)si268_e. The given atomic weight is an averaged value. Other prominent examples of this are lithium, boron, or chlorine.

Educt, Reactant, Product, Reagent. All of chemistry relies on the concerted rearrangement of atoms, the removal of bonds, or the reintroduction of new bonds. This is done by chemical reactions. A chemical reaction is referred to as the process that changes the atom arrangement and the bonds of a given molecule supplied to the reaction, creating a molecule of a different layout and (most likely) different chemical and physical properties. The material supplied to a chemical reaction as starting material is generally referred to as educt. The term reactant is used synonymously for educt. A chemical reaction often requires more than one educt to proceed. The rearranged molecule resulting from the reaction is referred to as product. If referring to either educt or product, the term reagent is used.

Stoichiometry. As stated, chemical reactions consume educts to form products. The nature of the chemical reaction at hand demands that certain quantities of educts are available and that a given quantity of product is created from them. The mathematical relationship between the amount of product and the amount of required educts is referred to as stoichiometry. We have already discussed the example of the creation of carbon dioxide from carbon and oxygen. Another example is the creation of magnesium oxide from magnesium and oxygen. The reaction is

2Mg+O22MgO

si23_e  (Rct 6.1)

The stoichiometry of this reaction implies that two molar parts of magnesium must be provided and one molar part of oxygen gas. If less than the required amount is given, which is referred to as under-stoichiometry, product is only produced until all the amount of the under-stoichiometrically provided educt is consumed. If “not enough” of one educt is provided, the second educt is provided in “too high quantity”, which is referred to as over-stoichiometry. Again, the limiting factor will be the educt provided under-stoichiometrically.

Equivalent. We have already seen that in some cases, a molecule contains more than one atom that may participate in a reaction. Let us return to the example of magnesium oxide that is created by burning magnesium (see Rct 6.1). We have corrected the stoichiometry by providing 2 molar equivalents of magnesium and only 1 molar equivalent of oxygen gas. We could also have written

Mg+12O2MgO

si24_e

in which case we have already attributed that 1 mol oxygen gas is able to provide two equivalents for the reaction. The feq for oxygen gas in this reaction is feq,o2=12si25_e. Equivalent factors are thus calculated by taking the reciprocal of the number of reactions one molecule can perform. Equivalents are important, e.g., for neutralization reactions using acids that provide more than one proton. Examples are sulfuric acid, which can provide two protons and thus has feq,H2SO4=12si26_e or phosphoric acid that has an equivalent of feq,H3PO4=13si27_e.

Concentrations. Often substances are dissolved in liquids for and during chemical reactions. The properties of these solutions change depending on the amount of substance. A good example are electrolytes which are usually made by dissolving substances which dissociate into their respective ions. There is more than one way to display concentrations. Usually if a specific substance is referred to (and the mixture contains more than one) an index will be given to the respective symbol, e.g., ci when referring to the c of substance i in a mixture. In general the solvent in which the substance is dissolved must be indicated, e.g., “in toluene”. If this is not the case, water is usually assumed to be the solvent.

Molar Concentration. The molar concentration is defined as the amount of substance per volume of the solvent

c=nVmolL1

si28_e

Very often molar concentrations are given with a capital M, e.g., 1 HCl that is equivalent to 1 mol L−1.

Mass Concentration. The mass concentration is defined as the mass of a substance per volume of the solvent

ρ=mVgL1

si29_e

Molality. The molality is defined as the amount of substance per mass of the solvent

b=nmmolkg1

si30_e

As an example, if an acid concentration is given by bHCl = 0.1mol kg−1 it contains 0.1 mol hydrogen chloride which is dissolved in 1 kg water. Molality refers only to the mass used and is therefore practical because it does not depend on thermodynamic conditions. The volume of the solvent is dependent on the temperature, the mass is not.

Mass Fraction. In a mixture, the mass fraction is defined as the mass of a given substance per overall mass of the mixture

wi=mimtotal=mij=1Nmjkgkg1

si31_e

As an example, the concentration of an aqueous acid is given as wHCl = 9 %. This means that 100 g of this acid will contain 9 g hydrogen chloride and 91 g water.

Mole Fraction. In a mixture, the mole fraction (or molar fraction) is defined as the amount of a given substance per overall amount of the mixture

xi=nintotal=nij=1Nnjmolmol1

si32_e

As an example, the concentration of an aqueous acid is given as xHCl = 9 %. This means that 100 mol of this acid will contain 9 mol hydrogen chloride and 91 mol water.

Equivalent Molar Concentration. The equivalent concentration is defined as the fraction of the molar concentration and the equivalence factor

ceq=cfeqmolL1

si33_e

In order to achieve complete neutralization of an acid, the equivalent concentrations of acid and base have to be identical. An example would be, if we want to neutralize a 0.1 mol hydrogen chloride solution with a dilute solution of sodium carbonate. The equivalent factors are feq,HCl = 1 and feq,Na2CO3=12si34_e. In order to neutralize completely, we have to use the same volume of a 0.1 mol hydrogen chloride and a 0.05 sodium carbonate solution.

Equivalent Amount of Substance. We have already used the equivalent amount of substance indirectly in the last example. It is given as the fraction of the amount of substance with the equivalence factor

neq=nfeqmol

si35_e

6.3 Important Terms and Concepts in Thermodynamics

Density. This equation is well-known but used often in thermodynamics, therefore we will shortly introduce the density to be

ρ=mV

si36_e  (Eq. 6.3)

This equation allows simplification of the specific volume (volume normalized to mass) to be simply expressed as

v=Vm=1ρ

si37_e

We will use this function when introducing a specific variable, i.e., thermodynamic variables normalized to the mass.

Control Volume. Thermodynamics uses control volumes in order to study changes of the important thermodynamic variables, e.g., the temperature. Often, a process is described conveniently if a suitable control volume is chosen. It requires some expertise and experience to choose control volumes correctly. For doing so, the system is enclosed with an imaginary boundary and the in- and outflux of mass and energy are controlled. The thermodynamic state variables of the system (i.e., the control volume) will change as a consequence of this in- and outflux across the boundary of the control volume. This in- and outflux brings in (or removes) energy from the system (see Fig. 6.2).

f06-02-9781455731411
Fig. 6.2 Schematic of a thermodynamic system as a control volume to and from which heat, work, and mass flows.

Thermodynamic System. In general, a thermodynamic system is referred to as a control volume for which the thermodynamic state variables (which we will introduce in a moment) are being studied.

Thermodynamic (System) State. The state of a thermodynamic system is defined by the current thermodynamic state variables, i.e., their values. Changes of states imply changes in the thermodynamic state variables. It should be noted that it is not important for a thermodynamic system by which processes the state variables were modified to reach their respective values.

Thermodynamic Process. In thermodynamics, a process is referred to as the transformation of a thermodynamic system from an initial thermodynamic state (which is described by the thermodynamic state variables) to a second thermodynamic state. The changes induced can be read from the change of the thermodynamic state variables.

Thermodynamics State Variables. The following lists the important thermodynamic state variables.

 pressure, denoted p and measured in Pa

 volume, denoted V and measured in m3; specific volume (normalized to the mass), denoted v and measured in m3 kg−1; please note that due to the fact that ρ=mVsi36_e and v=Vmsi39_e the specific volume is equal to 1ρsi40_e

 absolute temperature, denoted T measured in K

 heat, denoted Q and measured in J; specific heat (normalized to mass), denoted q and measured in J kg−1

 work, denoted W and measured in J; specific work (normalized to mass), denoted w and measured in J kg−1

 enthalpy, denoted H and measured in J; specific enthalpy (normalized to mass), denoted h and measured in J kg−1

 entropy, denoted S and measured in J K−1; specific entropy (normalized to mass), denoted s and measured in J K−1 kg−1

These variables are classified as either being intensive, that means they do not scale with the amount of substance contained in the system (temperature being a typical example), or extensive that means they scale with the amount of substance contained in the system (volume being a typical example). Some of these variables exist as specific variables which means they are normalized to the mass. These variables are usually denoted by lower case letters, e.g., h for specific enthalpy, s for specific entropy, or v for specific volume. Furthermore, variables may sometimes be normalized to a specific time, especially if dynamic processes are to be considered. These variables are then superscripted with a dot:

 mass flow (mass per time), denoted m˙=dmdtsi41_e and measured in kg s−1

 heat flow (heat per time), denoted Q˙=dQdtsi42_e and measured in J s−1; specific heat flow (heat flow normalized to mass), denoted q˙=dqdtsi43_e and measured in J kg−1 s−1

 work flow (work per time), denoted W˙=dWdtsi44_e and measured in J s−1; specific work flow (work flow normalized to mass), denoted w˙=dwdtsi45_e and measured in J kg−1 s−1

 enthalpy flow (enthalpy per time), denoted H˙=dHdtsi46_e and measured in J s−1; specific enthalpy flow (enthalpy flow normalized to mass), denoted h˙=dhdtsi47_e and measured in J kg−1 s−1

 entropy flow (entropy per time), denoted S˙=dSdtsi48_e and measured in J K−1 s−1; specific entropy flow (entropy flow normalized to mass) denoted s˙=dsdtsi49_e and measured in J kg−1 s−1

6.4 Ideal Gases

6.4.1 Definition

Most of the theoretical assumptions on gases will be made using the ideal gas equations. The concept of ideal gases is a simplification of the dynamics of real gases, but in most cases it is sufficient for describing a number of properties of real gases. Two basic assumptions are made for ideal gases.

 There are no attractive forces between the gas molecules - the only interaction of the gas molecules are collisions among the molecules and with the container wall.

 The total volume of all gas molecules is small compared to the total volume of the container in which the gas is stored - the gas molecules can thus be assumed as infinitesimal small points.

One of the known consequences of these assumptions is the fact that ideal gases fill the entire volume with which they are provided. If the volume is decreased, the pressure of the gas molecules increases as they are more tightly packed. If the temperature is increased without increasing the volume, the pressure increases. From these basic correlations, the ideal gas law is defined as

pV=nRT

si50_e  (Eq. 6.4)

pV=nNARNAT

si51_e  (Eq. 6.5)

pV=NkBT

si52_e  (Eq. 6.6)

p=nDkBT

si53_e  (Eq. 6.7)

R is the universal gas constant and has a value of 8.314 462 1 J mol−1 K−1 according to National Institute of Standards and Technology (NIST). Dividing R by NA results in the Boltzmann constant kB. Dividing the number of molecules by the volume results in the number density nD, i.e., the number of molecules per volume.

The ideal gas law (see Eq. 6.4) is often used in slightly modified form using Eq. 6.2 and Eq. 6.3 as

pV=mMRTpVm=RMTpρ=RSTp=ρRST

si54_e  (Eq. 6.8)

Here we have introduced the specific gas constant RS normalized to the molecular weight of the gas used. This is also a constant, although it is (obviously) different for every gas. A selection of commonly used gases is given in Tab. 6.1.

Tab. 6.1

Specific gas constants of commonly used gases [3]

GasMolar weight M g mol−1Specific gas constant RS J kg−1 mol−1
air28.970.287
CO244.010.189
CO28.010.297
H22.014.124
N228.010.297
O232.000.259
H2O, vapor18.010.462

Molar Volume. The volume any gas at ambient pressure will occupy can be calculated simply from the ideal gas law in Eq. 6.4. Using p = 1 bar = 105 Pa, n = 1 mol, R = 8.314 462 1 J mol−1 K−1 and a temperature of T = 0 si183_eC = 273.1499 K will result in a total volume of VM,0 = 0.022 414 m3 or 22.414 L. This volume L is commonly referred to as the molar volume. It is the volume occupied by 1 mol of a gas at standard conditions. It is equal for all gases. The number of molecules contained in this volume is referred to as the Loschmidt number NL. It is often confused with the Avogadro number NA, which defines the number of molecules per mole. The two numbers are interconnected by the molar volume

L=NANL

si55_e

The Committee on Data for Science and Technology (CODATA) recommends a value of 2.687 × 1025 for the Loschmidt constant. In general most gases, especially at atmospheric (or lower) pressure as well as elevated temperatures (room temperature or above), will be described sufficiently precise with the ideal gas law.

6.4.2 Mean Free Path

The most important interaction mechanism among the molecules of gases are collision events. The average length a gas molecule can travel before it encounters a collision event is referred to as the mean free path. We assume a particle of diameter d traveling at a velocity of v. In the time span t it will have traveled the distance

lt=vt

si56_e

Its cross-section area is given by πd2. Therefore it will have covered a total interaction volume Vv of

Vv=ltπd2=vtπd2

si57_e

in the time span t. The shape of this volume is cylindrical. Now, we assume that the particle is not alone in the control volume. We assume a total number of nD particles per unit volume (referred to as number density) to be in the control volume. The number of particles in the volume through which our particle traveled is given by Vv nD. Our particle would have collided with each of them. Therefore, the average mean path, i.e., the path traveled per collision would be

λ=distancetravelednumberofcollisions=ltVvnD=vtvtπd2nD=1πd2nD

si58_e  (Eq. 6.9)

Now this is only an approximation as it assumes the particles we collide with to be static. This is obviously not true. If our particle is moving, the other particles will move as well. Therefore, we need to take into account their relative velocity vdiffsi269_e instead of the absolute velocity. This velocity is given as

vdiff=(v1v2)2=v1v1¯2v1v2¯+v2v2¯

si59_e

where we consider the average velocities only. Now as v1si60_e and v2si61_e are uncorrelated and random, the average product v1v2¯si62_e will be zero as we do not expect the particles to move in a predominantly parallel or coordinated fashion. Furthermore they are of the same magnitude. Therefore we find

vdiff=(v1)2+(v2)2=(2v1)2=2v1

si63_e

which tells us that the average velocity of the particles will be higher by a factor of 2si64_e when considering collision with moving particles instead of collision with static particles. The average mean free path in Eq. 6.9 will therefore be smaller by this factor resulting in

λ=12πd2nD

si65_e  (Eq. 6.10)

Now the number of particles per volume nD can be calculated from Eq. 6.7 as

nD=pkBT

si66_e

in which we find Eq. 6.10 to be

λ=kBT2πd2p

si67_e  (Eq. 6.11)

For an example of how to approximate the diameter of a molecule see section 9.3.4. Some experimental values for commonly used gases are given in Tab. 6.2.

Tab. 6.2

Gas molecule diameters of some commonly used gases [4]

GasDiameter d mAverage velocityv¯meansi1_e m s−1
air0.36467
Ar0.36182
CO20.45379
H20.271 769
He0.221 256
Kr0.41274
N20.37475
O20.36444
Xe0.48219

6.4.3 Gas Pressure

We can estimate the gas pressure from kinetic theory by a similar simple model. Assume a particle being enclosed in a volume of width L along the x-axis. The particle moves in this container and collides with the side wall transferring the momentum

Δp=px(px)=2px=2mvx

si68_e

The collisions on one specific side of the container occurs every

Δt=2Lvx

si69_e

The force during this impact is given by

F=ΔpΔt=mvx2L

si70_e

If we assume a total of N number of gas molecules to be present in the container, we take the average of their velocities along the x-axis vx¯si71_e. The forces exerted by all of these molecules on one side of the container will then be

FN=Nmv¯x2L

si72_e

If we consider the average speed of all molecules in the container v¯2si73_e, we can again assume that their movement is random and not preferably in one direction. Therefore we can assume that the total fraction of the velocity along the x-axis will be a third of v¯2si73_e in which case we find

FN=Nmv¯23L

si75_e

Referring to the total force exerted onto the area L2 we find for the pressure

p=FNL2=Nmv¯23L3=Nmv¯23V=nDmv¯23

si76_e  (Eq. 6.12)

where we have used V = L3 to express the volume and nD as the number density, i.e., the number of molecules per volume.

6.4.4 Kinetic Energy

Next we will derive the kinetic energy, i.e., the correlation between the temperature and the movement of the particles. Applying the ideal gas law Eq. 6.6 to Eq. 6.12 we find

nDmv¯2V3=NkBT=pVmv¯23=kBT

si77_e  (Eq. 6.13)

Using the kinetic energy relation Ekin=12mv2si78_e for a single particle we can rewrite Eq. 6.13 as

Ekin=12mv2=32kBT

si79_e  (Eq. 6.14)

Obviously the total kinetic energy of the system is simply

Ekin,total=N12mv2=N32kBT

si80_e  (Eq. 6.15)

where we need to use to root mean square velocity, i.e., the average velocity of the particles.

Mean Square Velocity. Extending these models we can also derive the average speed of one molecule or particle usually as the root mean square velocity. It can be directly derived from Eq. 6.14 as

v=3kBTm

si81_e  (Eq. 6.16)

It gives the average velocity of all molecules in the control volume. Obviously, not all particles have the same speed. The speed follows the Maxwell1speed distribution. In the following section we will shortly derive this distribution.

6.4.5 One-Dimensional Velocity Probability Distribution

We begin by studying the velocity distribution along one axis. As we will see in a moment, this distribution is, mathematically speaking, a Boltzmann1distribution. We will derive the velocity probability distribution F (vx, vy, vz) which will give us the likeliness that a particle will travel at a given set of values for the independent variables vx, vy, and vz. So for each values of vx, vy, and vz, this distribution will give us the likeliness that a particle will travel at exactly these speeds. This value is 0 ≤ F (vx, vy, vz) 1 with the overall integral amounting to 1. This is due to the fact that each particle must travel at one set of velocities and if we integrate over all of them we will eventually catch all particles. As the movement of the particles in uncorrelated, i.e., it is random, we can surely assume that this combined three-dimensional distribution can be split into individual distributions along each axis that are independent. We therefore note

F(vx,vy,vz)=f(vx)g(vy)h(vz)

si82_e  (Eq. 6.17)

with f (vx), g (vy), and h (vz) being the velocity probability distributions along each axis. Again, these functions will return the likeliness of a particle traveling at a given velocity. We can already note here that these distributions will most likely look identical. This is due to the fact that the particle movement is random and we do not expect to see any preferred movement direction, and therefore no predominant velocity distribution. We quickly transfer this to a sum by taking the natural logarithm of Eq. 6.17 resulting in

ln(F(vx,vy,vz))=ln(f(vx))+ln(f(vy))+ln(f(vz))

si83_e  (Eq. 6.18)

One may wonder why we do this. If we assume the distributions to look the same in all directions, we may just as well derive only the distribution in one axis and simply repeat this process with the other two. Therefore, we now take the partial derivative with respect to vx (we could also take one of the other two) which results in

vxln(F(vx,vy,vz))=vxln(f(vx))+vxln(f(vy))+vxln(f(vz))=ddvxlnf

si84_e  (Eq. 6.19)

Now it becomes obvious why the natural logarithm helped: The distributions f (vy) and f (vz) have disappeared entirely because we converted the product of Eq. 6.4 to a sum (Eq. 6.18). We know that v2=vx2+vy2+vz2si85_e in which case we can rewrite Eq. 6.19 as

vxlnF=ddvxlnfvlnFvvx=ddvxlnf

si86_e  (Eq. 6.20)

We have to find vvxsi87_e which we find as

vvxv=vxvx2+vy2+vz2=12vx2+vy2+vz22vx=vxv

si88_e

that after inserting into Eq. 6.20 yields

vlnF1v=1vxdlnfdvx

si89_e  (Eq. 6.21)

As already stated, we could repeat the same process for the other two velocity components and this would result in a similar equation as Eq. 6.21 with only the right-hand side being different

vlnF1v=1vydlngdvy

si90_e  (Eq. 6.22)

vlnF1v=1vzdlnhdvz

si91_e  (Eq. 6.23)

Here we employ a trick we use intensively during the derivation of solutions to partial differential equations: If several differential equations all amount to the same value (in this case the left-hand side of Eq. 6.21), they must be constants (see section 8.3.3 for details on this neat little trick). We call this constant −λ and proceed with solving Eq. 6.21. Please note that the other two equations (Eq. 6.22 and Eq. 6.23) are solved likewise.

1vxdlnfdvx=λdlnfdvx=λvxdlnf=λvxdvxlnf=12λvx2+cf(vx)=c1e12λvx2

si92_e  (Eq. 6.24)

where we still need to find the two constants c1 and λ.

Determining c1. The c1 can easily be found by taking into account that the integral of the distribution must be 1, i.e., if we sum up all probabilities for all velocities, we must inevitably collect all particles. Therefore

+f(vx)dvx=+c1e12λvx2dvx=c1+e12λvx2dvx=!1

si93_e

The integral +eax2dxsi94_e is one of the Gaussian integrals and it evaluates to πasi95_e. You can also use Maple to verify that this is correct. In this case we find

c12πλ=!1c1=λ2π

si96_e  (Eq. 6.25)

Determining λ. Next we need to determine λ. This we do by taking into account that all kinetic energies of all particles combined must be identical to the overall energy of the system. This energy is given by Eq. 6.15 as

13N32kBT=12NkBT

si97_e  (Eq. 6.26)

The reason for the 13si98_e prefactor arises from the fact that we only take one third of the energy contribution as we are only considering the velocity component vx currently. Again, as the movement is random and uncorrelated, the energy contributions are identical. Now we know that the kinetic energy along the x-axis of one single particle is

Ekin=12mvx2

si99_e

We use Eq. 6.25 and integrate over all velocities in which case we find

Ekin,total=N12m+f(vx)vx2dvx=N12mλ2π+eλ2vx2dvx

si100_e  (Eq. 6.27)

where we have yet another Gaussian integral which evaluates to

+eax2x2dx=12πa32

si101_e

Therefore we can evaluate Eq. 6.27 and set it equal to Eq. 6.26 finding

Ekin, total=14Nmλ2ππ(λ2)32=12Nmλ=!12NkBTmλ=kBTλ=mkBT

si102_e

One-Dimensional Energy Distribution. Now we have all constants and can rewrite Eq. 6.24 as

f(vx)=m2πkBTemvx22kBT

si103_e  (Eq. 6.28)

This is the one-dimensional energy distribution, i.e., the probability function that a given particle will travel at a given speed which is equivalent to stating that a particle will have a certain kinetic energy. As already stated, this function is a Boltzmann distribution. We can quickly transfer it to a three-dimensional case.

Three-Dimensional Energy Distribution. Returning to Eq. 6.17 where we stated that the three energy distributions for vx, vy, and vz can be derived independently, we already stated that the functions we derive for each of them look identical (see Eq. 6.21, Eq. 6.22, and Eq. 6.23). Therefore we can reassemble the three-dimensional velocity probability distribution or energy probability distribution using Eq. 6.28 as a template resulting in

F(vx,vy,vz)=f(vx)g(vy)h(vz)=(m2πkBT)3(emvx22kBTemvy22kBTemvz22kBT)=(m2πkBT)32em2kBT(vx2+vy2+vz2)=(m2πkBT)32emv22kBT

si104_e  (Eq. 6.29)

Eq. 6.29 is the energy distribution in three dimensions.

6.4.6 Maxwell Speed Distribution

We now want to transfer the three-dimensional energy or velocity distribution into a function of only one variable. Instead of supplying the values vx, vy, and vz we want to supply only one value v thereby reducing the number of independent variables to only one. This is actually very simple to do: Eq. 6.29 gives us the probability of a particle traveling at a given set of velocities for vx, vy, and vz, we simply need to integrate over the individual independent variables. We therefore transfer F (vx, vy, vz) to spherical coordinates using one of the unit vectors as the velocity component. We therefore end up with a function F (θ, ϕ, v) which we simply need to integrate over the first two independent variables to result in a function of only the velocity as independent variable. In transferring the function to spherical coordinates, we have to take into account the Jacobian determinant (see section 7.5) in which case our integral has to be transformed as

dvxdvydvz=v2sinθdθdφdvF(vx,vy,vz)dvxdvydvz=F(θ,φ,v)v2sinθdθdφdv

si105_e

As stated, we do not want to integrate over the velocity here, therefore we only integrate twice. As usual, we integrate 0 ≤ θ ≤ π and 0 ≤ ϕ ≤ 2π in spherical coordinates. Thus we find

F(v)=0π02πF(θ,φ,v)v2sinθdθdφ=0πsinθdθ02πdφF(θ,φ,v)v2=0πsinθdθ02πdφF(θ,φ,v)v2=4πF(θ,φ,v)v2=4πv2(m2πkBT)32emv22kBT

si106_e  (Eq. 6.30)

Do not be confused because we are inserting v2 instead of r2 - we are working with velocities as the unit vector of our coordinate systems, which is why we have to use v instead of the parameter r which usually only indicates a length. Eq. 6.30 is the Maxwell velocity or speed distribution. It will return the likelihood of a particle to travel at the given velocity v. The unit of the function is 1ms1si107_e.

The distribution shown in Fig. 6.3 is calculated for a water molecule with M = 18 u at various temperatures. As you can see the strong shoulder observed at lower temperatures flattens out as the temperature increases. The whole distribution shifts toward the right, allowing a wider range of velocities. This is a very characteristic feature of gases at higher temperature.

f06-03-9781455731411
Fig. 6.3 Maxwell speed distribution calculated for a molecule of water with M = 18u at different temperatures. You can see the characteristic shoulder of this distribution moving toward the right as the temperature increases.

6.4.6.1 Mean Free Velocity

Eq. 6.30 is a very important function that allows us to calculate a number of important parameters. First of, we will derive the mean free velocity, which is the mean velocity obtained by integrating Eq. 6.30 over dv. We have already calculated the mean square velocity (Eq. 6.16) which gave us a rough estimate of the velocity of a free moving particle. The velocity we derive from the integral of Eq. 6.30 is more exact. It is given as

+vF(v)dv=+v4πv2(m2πkBT)32emv22kBTdv=4π(m2πkBT)32+v3emv22kBTdv

si108_e

This integral is best found with the help of an algebra tool, e.g., Maple which tells us that +x3eax2=12a2si270_e in which case we find

+vF(v)=4π(m2πkBT)32(2kBTm)2=8kBTπm

si109_e  (Eq. 6.31)

Eq. 6.31 is the mean free velocity. As you can see, it is close to the mean square velocity given by Eq. 6.16 which uses 3si110_e instead of 8si111_e although both values are reasonably close to each other.

6.4.6.2 Viscosity

The viscosity of a gas can also be estimated using the diameter of the gas molecules. It is given by [4] as

η=2.67×1020MTd2

si112_e

6.4.6.3 Partial Pressure

If gas mixtures are studied, the mixture, as well as each individual gas, is considered as a perfect gas. Therefore all individual gases will fill the entire volume. They will contribute to the overall pressure proportionally to their mole fraction

pAV=nARTpBV=nBRT

si113_e

with

p=pA+pBn=nA+nB

si114_e

Therefore the mole fraction nA is always equivalent to the partial pressure pA of a gas in a mixture.

Partial Pressure. If adding 1 L of nitrogen gas at a partial pressure of 0.8 bar and 1 L of oxygen gas at a partial pressure of 0.3 bar to a vessel of 1 L, the overall pressure inside of the vessel will be 1.1 bar. As a second example, ambient air has a composition of about 21 V% oxygen gas and 79 V% nitrogen gas. Thus the partial pressure of oxygen gas is pO2 = 0.21 bar and the partial pressure of nitrogen gas is pN2 = 0.79 bar.

6.5 Idealized Thermodynamic Processes

Thermodynamic processes can sometimes be idealized. During such a process, at least one of the thermodynamic state variables is assumed not to change. This often simplifies the description of a process significantly and is often a good approximation of the reality. The following idealizations are generally used

 isentropic - no change of entropy dSd=0si115_e

 isobar - no change of pressure dpd=0si116_e

 isotherm - no change of temperature dTd=0si117_e

 isochoric - no change of volume dVd=0si118_e

 adiabatic - no change of heat dQd=0si119_e

 stationary - no change of any of the thermodynamic state variables over time ddt=0si120_e

 closed - no in- or outflux of mass over time therefor no change of mass over time dmdt=0si121_e

 contained - no change of composition of the system over time, i.e., due to chemical reactions among individual compounds

There are several examples of such simplifications in reality. An example for isochoric processes are the change of pressure of a pressurized air bottle as the temperature increases. An example of a closed adiabatic system is a heat-insulated thermos flask. Additionally there are several structures also commonly used in microfluidics which do not change some of the state variables

 passive channel structures - do not exert work dWd=0si122_e

 nozzles - do not exert work dWd=0si122_e

 diffusers - do not exert work dWd=0si122_e

Reversible Processes. In thermodynamics, a process is considered reversible if the transition from one thermodynamic state to another can be reversed. Such transitions are rare in practice as it involves only equilibrium changes of the state variables, i.e., very slow transitions. Analytically, a thermodynamic process is reversible if the change of entropy (see section 6.7.1) is zero. Only then is it possible to reverse a change of state without loss of energy.

6.6 First Law of Thermodynamics

As stated, thermodynamic systems can be described using suitable control volumes (see Fig. 6.2). For such control volumes, the first law of thermodynamics states that the change of the total energy of the system is equal to the sum of all energy transferred into and out of the system

ΔEtot=d(imi(ui+epot,i+ekin,i))dt=jQ˙j+kW˙k+lm˙l(h1+vl22+ghl)

si125_e  (Eq. 6.32)

As can be seen, the sum of the total energy of the system Etot comprises the sum of all compounds (indexed with i) contained in the system. Each compound of mass mi contributes with its own inner energy ui and its own potential epot,i, and kinetic energy ekin,i. In addition to energy in- and outflux via transferred mass, energy in the form of heat and work may be transferred into and out of the system which is denoted with Qj.si271_e and Wk.si272_e, respectively. The first law of thermodynamics states that energy is never lost, it may be transferred out of the system, i.e., across the boundary of the control volume and into a second control volume, but it is not destroyed or lost (see Fig. 6.2).

In the following, we will look a bit closer into the individual components of the energy terms and what they relate to.

6.6.1 Work

The easiest component is work. It is the sum of all the volume forces acting on the control volume. This work is expressed as

dW=pdV

si126_e  (Eq. 6.33)

We will see a good example of such volume work in Fig. 6.4c.

f06-04-9781455731411
Fig. 6.4 Examples of irreversible (a) and reversible (b) processes. a) A gas volume is restricted to a small volume V1 by a removable separator. Removing this separator results in an irreversible expansion of the gas to the total vessel volume V2. b) If the restricted gas is heated, the separator is displaced and work can be gained from the system. The process is reversible. c) Work done by volume expansion as is the gas, e.g., in piston movement due to gas expansion.

6.6.2 Heat

Heat is the energy transferred into and out of the system at a given temperature. Combustion processes are typical examples of heat being applied to a system.

6.6.3 Internal Energy

The first energy term as seen from section 6.6 is the internal energy U of a thermodynamic system, which is defined as

U=Upot+Ukin

si127_e  (Eq. 6.34)

The energy is defined as the sum of all potential Upot and kinetic energy Ukin required in order to create the system. Ukin comprises the energy required for the motion of the atoms or compounds within the system. However, Ukin does not include the kinetic energy of the lateral movement of the system as a whole. Upot comprises the energy stored as mass (as well as by the electron configurations of the atoms), the chemical bonds (which may further react), and the physical forces (e.g., electromagnetic forces due to the presence of dipoles or internal stress/strain) within the system or among the atoms or compounds of the system.

6.6.4 Enthalpy

The internal energy is the energy required for creating a thermodynamic system. However, this system must occupy a given space and the energy required for creating this space needs to be added to the total energy required for creating and placing a system. This combined energy is referred to as enthalpy and is defined as

H=U+pV

si128_e  (Eq. 6.35)

The first term U sums up the internal energy of the system (see Eq. 6.34). The second term pV is the volume work done at the system or performed by the system.

6.6.5 Changes of Internal Energy and Enthalpy

Now if internal energy and enthalpy change over time, the total differentials of Eq. 6.34 and Eq. 6.35 have to be found. We will begin with the enthalpy

dH=dU+pdV+VdP

si129_e  (Eq. 6.36)

for which the last term is often neglected if processes at constant pressure are being studied. Often, the enthalpy has to be formulated as a function of pressure and temperature. Therefore, the partial differential is formulated as

dH=HTdT+Hpdp(realgas)

si130_e  (Eq. 6.37)

HTdT(idealgas)cpdT(perfactgas)

si131_e  (Eq. 6.38)

Here we have already introduced two simplifications when working with fluids in thermodynamics, namely the ideal gas and the perfect gas. In general, the latter term is often used interchangeably with the first. An ideal gas is a gas for which the enthalpy is only a function of the temperature. Whenever there is a linear correlation between enthalpy and temperature change (given by the proportionality constant cp) we refer to the gas as being a perfect gas. We have discussed the properties of ideal gases in section 6.4. In ideal gases, the enthalpy is only a function of the temperature, not of the pressure. In perfect gases (which are commonly used in thermodynamics), the dependence of the enthalpy on the temperature is a constant, the isobaric heat capacity cp.

In an analogy, the dependence of the internal energy is usually formulated as a function of the volume and the temperature. It is given by

dU=UTdT+UVdV(realgas)=UTdT(idealgas)=cvdT(perfectgas)

si132_e  (Eq. 6.39)

In ideal gases the change of the internal energy is only dependent on the temperature. For perfect gases this dependency is given by a constant, the isochoric heat capacity cv. The unit of this constant is J kg−1 K−1, but sometimes it is normalized to the molecular weight. The unit is then J g−1 K−1.

Heat Capacity of Fluids. Please note that for liquids (being incompressible), both enthalpy and internal energy are linearly dependent on temperature and the values of cp and cv can be assumed to be almost identical. For water, cv ≈ cp = 75.2724 kJ mol−1 = 4.1818 J g−1 K−1 if normalized to the molecular weight at standard ambient temperature and pressure (STP) conditions.

6.6.6 Reaction Enthalpy

The change of enthalpy is an important means of accessing chemical reactions. Consider Eq. 6.36 at constant pressure, which then simplifies to

dH=dU+pdV

si133_e

The amount of energy emitted from or absorbed by a system during a change of state is referred to as heat of reaction. The heat of reaction is characteristic for a chemical reaction. It is given in kJ, although sometimes the unit used is kcal where 1 kcal = 4.187 kJ. As it is dependent on the amount of substance used, it is normalized to 1 mol. The heat of reaction originating from 1 mol is referred to as the reaction enthalpyH.

Exothermic and Endothermic. Depending on whether or not a chemical reaction creates or absorbs heat, the reaction is either classified as being exothermic if heat is released or endothermic if heat is absorbed. By convention, for exothermic reactions ∆H < 0 whereas for endothermic reactions ∆H > 0.

A typical example of an exothermic reaction is the formation of ammonia from hydrogen gas and nitrogen gas

N2+3H2NH3ΔH=92.3kJmol1

si134_e  (Rct 6.2)

A typical example for an endothermic reaction is the formation of nitric oxide from oxygen gas and nitrogen gas

O2+N22NOΔH=+180.6kJmol1

si135_e  (Rct 6.3)

Temperature and Pressure Dependency. As we have seen during the study of the thermodynamics of the enthalpy, this value is dependent both on pressure and temperature. It is therefore required to always state at which conditions the respective reaction was carried out. Usually, the reaction will be done at STP conditions in which case the symbol ∆H0 is used and the value referred to as standard reaction enthalpy. It is usually sufficient to only indicate one ∆H value and the respective pressure and temperature as the ∆H for other pressure and temperatures can be calculated using Eq. 6.37 and the respective constants.

Hess’s Law. It is important to keep in mind that the thermodynamic state variables do not depend on the process used to obtain the thermodynamic system. This is referred to as Hess’s1law which states that by reacting educts to a given product each reaction path will add up to the same ∆H value. It is therefore not possible to find an “energetically cheaper” way to obtain a certain product.

As an example, we’ll study the reaction of carbon and oxygen gas to carbon dioxide. There are two distinct reaction pathways one can take. The first one is the direct oxidation given by

C+O2CO2ΔH0=393.8kJmol1

si136_e

The second option is the oxidation to carbon monoxide and a subsequent oxidation of carbon monoxide to carbon dioxide

C+12O2COΔH0=110.6kJmol1CO+12O2CO2ΔH0=283.2kJmol1

si137_e

If you add up the two ∆H values you will see that they are equal. Thus it is not important which reaction path is taken.

6.6.7 Standard Formation Enthalpy

Hess’s law gives rise to a very important concept in chemistry: The enthalpy change during a reaction can be calculated as the difference of “absolute” enthalpies of products and educts as

ΔHj0=HproductsHeducts

si138_e  (Eq. 6.40)

As enthalpies cannot be measured on an absolute scale, but only as changes of the enthalpy, the “absolute zero” value for the enthalpy is interpreted as the enthalpy that the most stable form, i.e., the most stable allotrope of an element would have at STP conditions. The entropy change that occurs while reacting this element to a given compound is referred to as the formation enthalpy. If the reaction is carried out at STP conditions, the enthalpy change is referred to as standard formation enthalpyΔHf0si2_e. Please note that this value refer to 1 mol of product. Taking our two earlier examples, we can give the standard formation enthalpy of ammonia to be ΔHf0(NH3)=46.1kJmol1si140_e and of carbon dioxide to be ΔHf0(CO2)=393.8kJmol1si141_e.

As an example of a more complex reaction, let us consider the following reaction which is the important reaction if the blast furnace process by which iron oxide iron(III)oxide (commonly referred to as rust) is converted to iron

Fe2O3+3CO(g)2Fe(s)+3CO2(g)

si142_e

for which the values for ΔHf0si2_e can be read from Tab. 6.3 as ΔHf0(CO2)=393.8kJmol1si141_e, ΔHf0(Fe2O3)=824.8kJmol1si145_e, ΔHf0(CO)=110.6kJmol1si146_e. For iron we find ΔHf0(Fe)=0kJmol1si147_e because it is an element. Therefore using Eq. 6.40 we obtain

Tab. 6.3

Standard formation enthalpy ΔHf0si2_e of some compounds. Taken from [8]

CompoundΔHf0si2_e, kJ mol−1CompoundΔHf0si2_e, kJ mol−1
gases
O3+142.8H2O-242.0
H2S-20.6CO-110.6
CO2-393.8SO2-297.0
liquids
HF-271.3HCl-92.4
HBr-36.4HI+26.5
H2O-286.0H2O2-187.9
NH3-46.1
solids
NaCl-411.3Fe2O3-824.8
Cgraphite0Cdiamond1.9
dissociation
H+218.1O+249.3
F+79.5Cl+121.8
N+472.5

t0020

ΔHf0(Fe2O3Fe)=0kJmol13393.8kJmol1(824.8kJmol13110.6kJmol1)=24.8kJmol1

si148_e

6.7 Second Law of Thermodynamics

6.7.1 Entropy

The first law of thermodynamics states that energy is conserved. This is the basis on which the enthalpy transitions of thermodynamic systems can be calculated. However, it does not state why and how physical and chemical transitions occur. This is what is described by the second law of thermodynamics. This law focuses on entropy, which was already introduced as a thermodynamic state variable.

The concept of entropy was originally introduced by Carnot.1 Entropy is sometimes referred to as the state of disorder of a system. However, definitions based on terms such as “disorder” or “chaos” are often ambiguous as the terms may have distinctly different meanings. Scientifically speaking, a system in equilibrium is in perfect disorder, meaning that there is no local gradient in concentration, temperature, etc., and therefore no useful (chemical or physical) potential from which energy could be gained. The entropy of such a system is high. Entropy may therefore, more intuitively, be defined as the degree of dispersal of energy or the absence of any gradient or potential, i.e., the homogeneity of the energy distribution. A good example is a block of ice placed in a warm atmosphere. Initially there is a large gradient in temperature which may be used to gain useful energy from the system. One example of energy usage of such as system would be an energy generator such as a Seebeck generator that exploits temperature gradients to create electrical potential differences and thus, current. Over time, the ice will melt and the temperatures will equilibrate. The system has increased its entropy as the temperature gradient, i.e., the temperature potential has diminished.

The second law of thermodynamics states that only thermodynamic transitions which increase the entropy will effectively happen. This is very important, e.g., for predicting if a certain chemical reaction will take place or not. If the reaction would result in the increase of entropy, then this reaction can (in principle) take place. Similar to enthalpy values, which can only be given as relative values and for which therefore an arbitrary zero value has to be defined, entropy values are also referenced. By definition, an ideal crystalline substance or a perfect gas at absolute zero temperature 0K will have an entropy of zero. As we will see, it is sufficient to have a reference entropy value from which the entropy at any given temperature and pressure can be calculated. The most commonly used entropy value is the standard entropy S0 which is the entropy of 1 mol substance at STP conditions. This value is a material constant (see Tab. 6.4). Looking at Tab. 6.4 some interesting details can be gained

Tab. 6.4

Standard entropy S0 of some compounds [8]

CompoundS0, J mol−1CompoundS0, J mol−1
gases
H2130.7F2202.8
Cl2223.1N2191.6
O2205.2O3239.0
I2260.8H2O188.8
CO197.7CO197.7
CO2213.8HF173.8
HCl186.9
liquids
H2O70.0
solids
Cgraphite5.7Cdiamond2.4
Fe27.3S31.8
Ca41.7CaO39.8
CaO39.8Fe2O387.5
I2166.2
dissociation
H114.7F158.8
Cl165.2N153.3
O161.1

t0025

 The most chemically stable compounds (such as graphite or diamond) have low S0 values.

 Solids have lower S0 values than liquids or gases.

 Liquids have lower S0 values than gases.

 Solids with a high degree of order (as is the case in crystals) have very low S0 values.

6.7.2 Entropy Changes due to Chemical Reaction

During a reaction, the entropy changes. Large changes of entropy occur if the state of a reagent is changed, e.g., if a liquid and a solid react to a gas

12O2(g)+C(s)CO(g)ΔS0=197.7(205.22+5.7)Jmol1=+89.4Jmol1

si149_e

Very small values are obtained if gases are converted to liquids or solids

Ca(s)+12O2(g)CaO(s)ΔS0=39.8(41.7+205.22)Jmol1=104.5Jmol1

si150_e

In general, the entropy change during a phase transition can be calculated as

ΔS=ΔHvapT10.5R85JK1mol1to88JK1mol1

si151_e  (Eq. 6.41)

This is the case as phase transitions are isothermal processes, i.e., the temperature of a medium does not change during a phase transition. As stated, this value is interestingly constant for many liquids and can be approximated by Eq. 6.41 at STP conditions. This approximation is referred to as Trouton’s1rule. Tab. 6.5 states some of the most relevant enthalpies of vaporization.

Tab. 6.5

Standard enthalpy of vaporization ∆Hvap of some compounds. Please note that the values for the given temperature may be different from the enthalpy of vaporization at the actual boiling point [4, 10]

CompoundTemperature °CHvap kJ mol−1
gases
ammonia2519.86
bromine gas2530.91
chlorine gas2517.65
hydrogen sulfide2514.08
carbon disulfide2527.51
liquids
water045.05
2543.99
4043.35
6042.48
8041.59
10040.66
12039.68
14038.64
16037.52
18036.30
20034.96
22033.47
24031.81
26029.93
28027.80
30025.30
32022.30
34018.50
36012.97
3742.07
hydrogen chloride259.08
thionyl chloride2531.00
nitric acid2539.10
hydrogen peroxide2551.60
chloromethane2531.28
nitromethane2538.27
1,1,2-trichloro-1,2,2-trifluorethane2528.08
acetonitrile2532.94
acetic acid2523.36
N,N-dimethylformamide2556.19
ethanol2542.32
dimethyl ether2518.51
acetone2530.99
isopropanol2545.39
toluene2538.01
o-xylene2543.43
m-xylene2542.65
p-xylene2542.40
anisole2546.90

t0030

6.7.3 Mathematical Description

Mathematically, the entropy for a reversible isothermal process is described as

dS=dQT

si152_e  (Eq. 6.42)

It has to be noted that most thermodynamic processes are not isothermal and thus the entropy is usually expressed in diagrams as function of the temperature S (T). Therefore according to Eq. 6.42, the heat transferred is the integral Q = ∫ S (T) dT.

Using Eq. 6.42 and Fig. 6.2 the second law of thermodynamics can be expressed in its most general form as

diSidt=kQ˙kTk+im˙lsl+S˙produced

si153_e  (Eq. 6.43)

where the individual terms describe the following contributions

 diSidtsi154_e describes the temporal changes of the entropy values Si of all compounds i contained in the system

 kQ˙kTksi155_e describes the entropy transferred into and out of the system by heat flows Q˙ksi156_e at constant temperatures Tk; this term is zero for adiabatic processes

 lm˙lslsi157_e describes the entropy brought into and out of the system by all mass flows m˙lsi158_e; this term is zero for closed and contained systems

 S˙producedsi159_e describes the entropy produced inside of the system during the respective process; this term is zero for isentropic processes

Please note that the overall change of entropy can never be negative, otherwise the respective process or reaction would not occur at all. Now the question arises, how can some chemical reactions have negative S0 values? The answer to this is that the overall process must have a positive change of entropy. If a single reaction S˙producedsi159_e within the control volume of this process has a negative change in entropy, the system would need to compensate this by providing heat Q˙kTksi161_e and thus raise the overall change of entropy to a positive value.

6.7.4 Reversible and Irreversible Processes

Let us study examples for reversible and irreversible processes (see Fig. 6.4). A gas volume is restricted inside of a vessel to a fraction V1 if the total volume by a separator which can be removed or displaced. Removing the separator will result in expansion of the gas to the total volume of the vessel (Fig. 6.4a). If the gas is heated, it will expand and displace the separator (see Fig. 6.4b). In both cases, we assume an isothermal adiabatic process. The system is closed and contained and we assume ideal gas behavior.

We will start discussing the case displayed in Fig. 6.4b. Taking Eq. 6.32, we can therefore assume

W+Q=d(mu)dt=mdudt

si162_e

As stated, we assume isothermal expansion and ideal gases therefore taking Eq. 6.39, we assume the overall change of internal energy to be zero. Therefore

Q=W

si163_e  (Eq. 6.44)

Furthermore, the overall entropy change of the system must be zero if the process is (as assumed) reversible. Taking Eq. 6.43 we derive

dSdt=Q˙T+S˙produced=0

si164_e  (Eq. 6.45)

Therefore the system will need to consume the same amount of entropy as provided by means of heat transfer. As entropy cannot be consumed, the ambiance has to provide this entropy. In consequence, the entropy of the gas will be increased by heat transfer, whereby the entropy of the ambient will be reduced by the same amount. The important thing to note is that the entropy provided by the ambiance can be restored if the system would provide heat to the ambient, thereby reversing the process.

The expansion can be used to do work W (see Fig. 6.4c). This work can be expressed as

dW=pAdx=pdV

si165_e

Using the ideal gas law Eq. 6.4 we derive

p=nRTV

si166_e

and therefore

dW=nRTVdVW=nRTlnV+c1

si167_e

W12=nRTlnV1V2

si168_e  (Eq. 6.46)

=nRTlnp2p1

si169_e  (Eq. 6.47)

=TΔS12

si170_e  (Eq. 6.48)

This is the volume work performed by the system upon expansion. It also gives the amount of energy required for achieving this expansion using Eq. 6.44, as well as the gain of entropy of the system and the reduction in entropy of the ambiance using Eq. 6.45.

We now consider the case displayed in Fig. 6.4a. In this case, there is no heat provided to the system, but the system will perform the same amount of work (see Eq. 6.45). However, using Eq. 6.45 we can see that the ambiance does not need to provide entropy in this case. Therefore the gain of entropy of the system is not balanced, it simply increases. This entropy corresponds to the total expansion work of the system, which is lost irreversibly.

6.7.5 Thermodynamic Equations of State for Ideal Gases

For ideal gases, the first and the second laws (see Eq. 6.43) of thermodynamics can be simplified. In the following we consider a closed and contained system in a static control volume, i.e., we neglected any contribution of potential or kinetic energy. The left-hand side of Eq. 6.32 can thus be simplified to

ΔEtot==dUdt

si171_e

Adding the right-hand side of Eq. 6.32 we obtain

dU=dQ+dW

si172_e

which we make specific by division by mass to read

du=dq+dWm

si173_e  (Eq. 6.49)

Using Eq. 6.33 we can express the work as

dWm=pdWm=pd1ρ=pρ2dρ

si174_e  (Eq. 6.50)

For the last equation we have used the chain rule of differentiation (see Eq. 3.14). Using Eq. 6.42 we can express the heat as

dp=Tds

si175_e  (Eq. 6.51)

Eq. 6.49 can be written using Eq. 6.50 and Eq. 6.50 as

du=Tds+pρ2dρ

si176_e  (Eq. 6.52)

From Eq. 6.35 we know that the enthalpy can be expressed as

H=u+pVm=u+pρ

si177_e  (Eq. 6.53)

In order to find the change of enthalpy, we need to find the total differential of Eq. 6.53 which is

dH=Hudu+Hpdp+Hρdρ=1du+1ρdp=pρ2dρ

si178_e  (Eq. 6.54)

We still need the total differential of the internal energy du which we have already derived in Eq. 6.52. Using this equation, Eq. 6.54 can be rewritten to

dH=Tds+pρ2dρ+1ρdppρ2dρ=Tds+1ρdp

si179_e  (Eq. 6.55)

From Eq. 6.55 the temperature and density of a gas can be calculated as

T=HT|p:=cp

si180_e  (Eq. 6.56)

1ρ=HT|V:=cv

si181_e  (Eq. 6.57)

Using Eq. 6.56 and Eq. 6.57 we can simplify Eq. 6.53 to

H=u+pρ=cpTu=Hpρ=cpTpρ=cvT

si182_e  (Eq. 6.58)

where we have used Eq. 6.38 and Eq. 6.39 which state that in an ideal gas cp and cv are constants. Eq. 6.58 is often referred to as the thermodynamic equation of state for an ideal gas. This equation of state allows us to express the internal energy as a function of pressure, density, and temperature.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset