6.7.6 Gibbs Free Energy

As we have seen, both the enthalpy H as well as the entropy S are important when describing thermodynamic processes. Now the question arises, how are they interdependent, i.e., how are they connected? This interdependency is described by the Gibbs free energy G. The Gibbs free energy is often referred to as free enthalpy or free enthalpy of reaction.

The Gibbs free energy is defined as

G=U+pVTS=HTS

si183_e  (Eq. 6.59)

As stated, the enthalpy is a function of temperature and pressure which makes G likewise dependent on pressure and temperature. If all reagents are provided at STP conditions and referenced to 1 mol product, the resulting Gibbs free energy is referred to as free standard enthalpy of reactionG0

ΔG0=ΔH0TΔS0

si184_e

If a chemical process is exothermic, work can be gained from the reaction. The maximum work gained is given by the changes in entropy, given by ∆H0 reduced by the contribution of the entropy. As stated, the entropy is the amount of work which cannot be harvested from the system. This is reflected by Eq. 6.59. In general, exothermic reactions have negative ∆G0 values, whereas endothermic reactions have positive ∆G0values. Endothermic reactions require energy from the outside in order to proceed. A reaction for which ∆G0 is almost zero is considered to be an equilibrium.

6.7.6.1 Gibbs Free Energy for Electrochemical Reactions

The Gibbs free energy is, of course, also defined for electrochemical reactions where it defines the electromotive force of the reaction

G=zFΔμ

si185_e

If all reagents are provided at STP conditions, ∆G0 is given as a function of the standard reduction potential E0

G=zFΔE0

si186_e

6.7.6.2 Free Standard Enthalpy of Formation

As the Gibbs free energy depends on the enthalpy, it also suffers from the fact that its values cannot be given in absolute quantity, but only relative to a reference value. As stated, the standard reaction enthalpy ∆H0 is referenced to 1 mol of substance at STP conditions while setting ∆H0 = 0 for the elements. This gives rise to the free standard enthalpy of formation ΔHf0si2_e. In an analogy, the Gibbs free energy G of the elements is set G = 0 and the reaction considered at STP conditions for 1 mol of product. This gives rise to the free standard enthalpy of formation

ΔGf0=ΔGfproducts0ΔGfeducts0

si188_e

Some of the most commonly used ΔGf0si5_e values are given in Tab. 6.6.

Tab. 6.6

Free standard enthalpy of formation ΔGf0si5_e of important compounds

CompoundΔGf0si5_e kJ mol−1CompoundΔGf0si5_e kJ mol−1
gases
O3+163.3H2O, gas-228.7
CO-137.2CO2-394.6
SO2-300.4NO86.6
liquids
H2O70.0HF-273.4
HCl-95.4HBr-53.5
HI1.7H2O, liquid-237.3
NH3-16.5
solids
NaCl-384.3CaO-604.6
Al2O3-1 583.5SiO2-857.3
CuO-129.8
dissociation
H+203.4O+231.9
F+62.0Cl+105.8
N+455.9

t0035

Free Standard Enthalpy of Formation of Some Reactions. With the values given in Tab. 6.6 the ΔGf0si5_e values of chemical reactions can be calculated. As an example, let us consider the following reaction

CO+12OCO2ΔGf0=ΔGf0(CO2)(12ΔGf0(O2)+ΔGf0(CO))=(394.6(0137.2))kJmol1=257.4kJmol1

si191_e

Likewise the reaction of carbon and oxygen gas to carbon dioxide is given by

C+O2CO2ΔGf0=ΔGf0(CO2)(ΔGf0(C)+ΔGf0(O2))=394.6kJmol1

si192_e

The reaction of hydrogen gas and chlorine gas is given by

12H2+12Cl2HClΔGf0=ΔGf0(HCl)(12ΔGf0(H2)+12ΔGf0(Cl2))=95.4kJmol1

si193_e

A reaction with a positive value for ΔGf0si5_e is the reaction of nitrogen gas and oxygen gas to nitric oxide

12N2+12O2NOΔGf0=ΔGf0(NO)(12ΔGf0(N2)+12ΔGf0(O2))=+86.6kJmol1

si195_e

6.7.6.3 Interpretation of the Gibbs Free Energy

The Gibbs free energy is an important means for predicting chemical reactions. As already stated, chemical reactions are either exothermic or endothermic, depending on whether or not they require energy from the outside in order to proceed. In adiabatic closed and contained systems, only exothermic reactions can occur as no heat can be transferred from the outside. Transferring heat into a system (endothermic reactions) reduces the entropy of the ambiance while increasing the entropy of the system. Transferring heat out of the system (exothermic reactions) reduces the entropy of the system while increasing the entropy of the ambiance. Therefore, looking at Eq. 6.59 the borderline case is the case for which G = 0 and therefore

ΔSsystem=ΔHTΔSambiance=ΔSsystem

si196_e

For an exothermic reaction ∆H < 0 and therefore the entropy change of the ambiance is positive as heat is transferred out of the system. For endothermic reactions ∆H > 0 and therefore the entropy change of the ambiance is negative as heat is transferred into the system. However, in chemical reactions, the entropy change due to the reaction must also be considered (see Eq. 6.43). Thus adding up the entropy contribution and the enthalpy change, three cases must be considered.

 ΔGf0<0si197_e: the reaction will occur spontaneously and work may be gained from it

 ΔGf0>0si198_e: the reaction will not occur spontaneously; if the entropy is increased (by external heating), the Gibbs free energy may become sufficiently negative for the reaction to occur

 ΔGf0=0si199_e: the reaction is in equilibrium; usually values between − 5 kJ mol−1 and + 5 kJ mol−1 are considered to be equivalent to 0

It is important to note that merely judging a reaction exothermic and endothermic by looking at the enthalpy changes alone is not a sufficient manner in which to judge if a reaction will effectively occur. As the changes in entropy (especially if the formation or consumption of a gas is involved) may be significant, exothermic reactions may still yield a positive Gibbs free energy and thus, will not occur spontaneously. Also, ΔGf0si5_e gives no information about the speed of the reaction, but merely about the degree of completion. Reactions with values ΔGf00si201_e will tend to run to completion while reactions with values ΔGf00si202_e will yield almost exclusively the educts. ΔGf0si5_e therefore only refers to the thermodynamic stability of a product that is equivalent to considering the equilibrium of a reaction. As an example, we have calculated in section 6.7.6.2 that the reaction of hydrogen gas and chlorine to hydrogen chloride has a negative value ΔGf0si5_e. However, this gas mixture is stable at ambient conditions. So thermodynamically, the reaction can occur (in principle) - however, it is so slow that it cannot be observed effectively. It is therefore important to note that the Gibbs free energy will indicate if a certain chemical reaction is thermodynamically possible. However, it gives no information on the speed of the reaction.

Determining the Gibbs energy of a substance allows characterizing the ability of this substance to further react. This is sometimes referred to as thermodynamic potential. A system with high potential is able to react further, and a system with low potential will most likely not be reactive. However, the fact that a certain state may be thermodynamically more stable than another does not allow us to directly deduce whether or not this transition really occurs and at what transfer rate. A good example for this is the transition between different states of carbon. Thermodynamically, the graphite state of carbon is more stable than the diamond state. However, at ambient conditions, the rate of this reaction is so low that it cannot be observed.

Diamonds Are Forever? Another interesting example of the interpretation of the Gibbs free energy is the conversion of diamond to graphite, both of which are allotropes. The Gibbs free energy in this case is

ΔGf0(CdiamondCgraphite)=ΔGf0(Cgraphite)ΔGf0(Cdiamond)T(ΔS0(Cgraphite)ΔS0(Cdiamond))=(0kJmol11.9kJmol1)298.15K(5.7kJmol12.4kJmol1)=2.88kJmol1

si205_e

As we can see, this value is negative, but not very much smaller than 0. So in general, this reaction can occur in principle thermodynamically but it is so slow that the conversion can be neglected. The reason why ΔGf0(Cdiamond)=0kJmol1si206_e and ΔHf0(Cdiamond)>0kJmol1si207_e is due to the fact that graphite is the more thermodynamically stable allotrope of carbon and therefore chosen to be zero for ΔHf0(Cgraphite)=0si208_e reference.

Qualitative Temperature Dependency. In a moment we will have a look at the pressure and temperature dependency of the Gibbs free energy. But we can already discuss the influences of ∆H0 and S0 qualitatively. We will briefly discuss the four possible cases.

 H0 > 0, S0< 0: the Gibbs free energy is positive in all cases, there will be no reaction even at elevated temperatures

 H0 < 0, S0 > 0: the Gibbs free energy is always zero; at lower temperatures the enthalpy change will dominate at higher temperatures the entropy change

 H0 < 0, S0 < 0: the Gibbs free energy is negative at lower temperatures, but will become positive if the temperature is chosen too high

 H0 > 0, S0 > 0: the Gibbs free energy is positive at lower temperatures dominated by the enthalpy; at higher temperatures these reactions can proceed due to the increase in dominance of the entropy

A common example is phase changes that can be characterized using the Gibbs free energy. Gases have higher entropy than liquids, which makes their Gibbs energy lower than that of solids. However, at lower temperatures, this term quickly declines and the solid state becomes thermodynamically more stable. This is why liquids and gases are solid at lower temperature.

Equilibrium Point of Water. Let us consider this as an example. Water at STP conditions in the gaseous state has ΔHf0(g)=242.0kJmol1si209_e and S0 (g) = 188.8 J mol-1. In the liquid state, it has ΔHf0(l)=286.0kJmol1si210_e and S0 (l) = 70.0 J mol−1. The values can be taken from Tab. 6.3 and Tab. 6.4. The equilibrium point is reached, if the two Gibbs free energies are equal

G(g)=ΔHf0(g)TbS0(g)=ΔHf0(1)TbS0(1)=G(1)

si211_e

From this equation results Tb = 370.37 K = 97.22 °C. This is a very good approximation for the actual boiling point of water, which is the condition at which liquid water is in equilibrium with water vapor. In this example we used the enthalpy and entropy values for STP conditions, for a more precise calculation, the values should have been corrected with increasing temperature.

6.7.6.4 Temperature Dependency of the Gibbs Energy

Before discussing the temperature and pressure dependence of the Gibbs free energy, we will introduce a couple of more important thermodynamic equations. The first equation helps us link the first law of thermodynamics to the second law of thermodynamics, which we have done already in section 6.7.4. When considering closed and contained systems, the heat flow Q.si273_e can be expressed using Eq. 6.43 as

dQ=TdS

si212_e  (Eq. 6.60)

Using Eq. 6.32 and Eq. 6.33 we can deduce

dW+dQ=dUpdV+TdS=dUpdV+TdS=d(HpV)pdV+TdS=dHpdVVdp

si213_e  (Eq. 6.61)

dH=TdS+Vdp

si214_e  (Eq. 6.62)

Both Eq. 6.61 and Eq. 6.62 are used often in thermodynamics and are commonly referred to as fundamental equations. We will also employ them during the derivation of the temperature and pressure dependence of the Gibbs free energy. Now we will write Eq. 6.59 as a differential of the thermodynamic state variables which yields

dG=dHd(TS)dH=dG+d(TS)

si215_e

Using Eq. 6.63 we obtain

TdS+Vdp=dG+d(TS)dG=VdpSdT

si216_e  (Eq. 6.63)

We are interested in the change of the Gibbs free energy over temperature for which we have to find the partial differential with respect to the temperature. We will assume the pressure to remain constant, i.e., the reaction is carried out in a sufficiently large vessel. We rearrange Eq. 6.59 to read

S=GHT

si217_e  (Eq. 6.64)

Using Eq. 6.63 we now derive GT|p=Ssi218_e. We will need this expression in a moment

GT|p=S

si218_e

This equates to Eq. 6.64 and thereby yields

GT|p=HTGT|pGT=HT

si220_e  (Eq. 6.65)

Now the left-hand side of this equation can also be written as TGT|psi221_e. This is not straightforward to see, but can easily be checked by applying Eq. 3.13

TGTT|p=T(1TGT|pGT2)=GT|pGT

si222_e

With this we can rewrite Eq. 6.65 as

GTT|p=HTGTT|p=HT2

si223_e  (Eq. 6.66)

This equation is referred to as the Gibbs-Helmholtz equation which gives the temperature dependency of the Gibbs energy as a function of the temperature at constant pressure. It was derived independently by Gibbs and German physicist Hermann von Helmholtz.1 Before solving this differential equation in the following section we will shortly formulate the pressure dependency of the Gibbs free energy.

6.7.6.5 Pressure Dependency of the Gibbs Free Energy

Now we will study the pressure dependency of the Gibbs energy at constant temperature. We employ Eq. 6.63 setting SdT = 0 as the temperature is constant yielding

dGdp|T=Vdp=nRTdpp

si224_e

Here we have used the ideal gas equation (Eq. 6.4). It is solved to

ΔG12=nRTlnp2p1

si225_e

This equation is used in order to calculate the Gibbs energy if the reagents are provided at different partial gas pressures, as the increase in Gibbs energy can simply be calculated as a thermodynamic process 1 → 2. So if the Gibbs energy at STP conditions, i.e., pressure p0 is given, the Gibbs energy at any pressure can be calculated as

Gp=ΔG0+ΔGp0p=ΔG0+nRTlnpp0

si226_e

If more than one reaction partner takes part in the reaction, the changes in Gibbs energy are summarized. If we now consider the arbitrary reaction

nA+ nBBnCC

si227_e

where the coefficients ni account for the stoichiometry of the reaction, the overall change of Gibbs free energy can be obtained as the sum of the changes of the individual compounds as

ΔGstartend=ΔG0+RT(nCln(pC,endpC,start)(nAln(pA,endpA,start)+nBln(pB,endpB,start)))=ΔG0+RT(ln(pC,endnCpA,startnApB,startnBpC,startnCpA,endnApB,endnB))=ΔG0+RTln(KpKp0)

si228_e  (Eq. 6.67)

where we have introduced the equilibrium constant of the reaction at a given pressure Kp and the equilibrium constant of the reaction at STP conditions Kp0si229_e. The equilibrium can also be described for solutions using the respective molar concentrations at STP conditions Kc and Kpcsi230_e.

If the reaction is carried out in equilibrium and in a very large vessel, the Gibbs free energy does not change and ∆Gstart→end = 0. In this case, ∆G0 can be calculated as

ΔG0=RTln(KpKp0)

si231_e  (Eq. 6.68)

Electrochemical Reactions. For electrochemical reactions, the dependency of the Gibbs energy on the concentrations is likewise described by

ΔGstartend=E0RTzFln(KpKc0)

si232_e

Van’t Hoff Equation. We now finish the formulation of the temperature dependency of the Gibbs free energy. We have already derived the Gibbs-Helmholtz equation (see Eq. 6.66) which we rewrite to

ddT(ΔG0)T=ΔH0T2

si233_e

where we use Eq. 6.68 for the left-hand side to obtain

ddTln(KpKp0)=ΔH0RT2

si234_e  (Eq. 6.69)

This equation is referred to as the Van’t Hoff1equation. It is an important equation in practical chemistry as it forms the basis of the Q10temperature coefficient which predicts that the reaction speed of a chemical reaction is approximately doubled (sometimes even tripled) if the temperature is raised by 10 K. This prediction is surprisingly correct for many systems in chemistry, biochemistry, and biology.

Equation Eq. 6.69 can be solved using Eq. 6.68 to result in

ln(KpKp0)=ΔH0R(1T1T0)

si235_e  (Eq. 6.70)

This equation gives the dependency of the reaction equilibrium with respect to temperature. If the equilibrium constant Kp1 at a temperature T1 is known, the equilibrium constant Kp2 at a temperature T2 can be calculated.

We have thereby established an analytical correlation between the Gibbs energy and the pressure as well as the temperature, respectively. As stated, the Gibbs energy is dependent on these parameters which we may change during the reaction. If changing these parameters would result in a shift of the Gibbs free energy, this may help optimizing reaction conditions.

6.8 Third Law of Thermodynamics

We have worked with the first and the second law of thermodynamics and we will briefly introduce the third law of thermodynamics as well. The third law of thermodynamics states, that the absolute zero-temperature 0 K cannot be reached by a finite number of steps. Therefore any technical reference to absolute zero is always referred to as an approximated value. Technically, temperatures as low as 100 pK have been obtained in the laboratory [12] and temperatures as low as 3 K observed experimentally in space [13]. For low temperature applications, usually liquid gases are used into which the respective experiment is immersed. Some of the most commonly used cooling gases are summarized in Tab. 6.7. Usually, noble gases such as argon, or helium, or atmospheric gases, mostly nitrogen gas, are used for this purpose as they are chemically stable and easy to handle. Using helium allows achieving temperatures down to − 268.93 °C which is close to 4.22 K.

Tab. 6.7

Boiling points of some commonly used gases for low-temperature applications [4]

CompoundBoiling point, °CComment
water+100fixed property
ammonia-33.33
carbon dioxide-56.56dry ice
oxygen gas-182.95
argon-185.85
nitrogen gas-195.79
hydrogen gas-252.87
helium-268.93

t0040

6.9 Heat and Mass Transfer

6.9.1 Introduction

Having reviewed the most important concepts of thermodynamics, we will now discuss two important concepts which are often considered alongside the fundamental concepts of thermodynamics. These concepts are heat transport and mass transport. These two mechanisms are among the most important transport phenomena.

Heat Transfer: Conduction.Heat transfer is the process by which a material transports heat through its bulk. Heat is a conservative quantity and can thus be transported. The phenomena is well known from everyday life. If we drop a metal spoon in a cup filled with a hot beverage, e.g., coffee, the spoon will heat up. This is due to the fact that heat is transferred from the hot fluid into the metal and the metal is able to conduct the heat all the way up to our fingers. Intuitively we know that some materials are better conductors for heat than others. So using a spoon which has a polymer handle instead of a solid-metal spoon will allow us to stir the hot beverage without the risk of burning our fingers.

Mass Transfer: Diffusion. Similar to heat transfer, mass transfer is a phenomenon we know from everyday life. If we drop a piece of sugar into our coffee we can see that the sugar will dissolve over time. Usually, we stir the coffee to facilitate mixing. However, even if we do not mix it manually, the coffee will become sweet because the sugar will diffuse over time. This transport of mass is mainly due to the fact that the concentration of sugar is non-homogeneous throughout the bulk of the liquid. Thus, gradients in concentration are a driving factor for material transport by diffusion. As we will see, the two main transport mechanisms for mass are convection (in our example by stirring) and diffusion.

6.9.2 Fourier’s Law of Heat Conduction

6.9.2.1 Stationary Heat Conduction

Fourier’s Law of Heat Conduction in One Dimension. Fourier’s law of heat conduction is named after French mathematician Jean-Baptiste Fourier. The history of this important fundamental equation of heat conduction has been reviewed by Narasimhan [14] in an article worth reading. Fourier observed that the heat transported through a thin layer of a material was proportional to the differences in temperature of the two sides of the layer (Fig. 6.5a)

f06-05-9781455731411
Fig. 6.5 Fourier’s law of heat conduction. a) Heat will propagate through a layer of material from the side with the higher temperature to the side with the lower temperature. The temperature profile which forms is often non-linear.

q˙A=kT2T1X=λT2T1X=λdTdx

si236_e  (Eq. 6.71)

As the specific heat q˙Asi239_e is proportional to the area across which the heat is transferred, it makes sense to refer the heat to the area therefore

dQ˙=q˙AdA

si237_e  (Eq. 6.72)

=λdTdxdA

si238_e  (Eq. 6.73)

The unit of q˙Asi239_e is therefore J m−2 s−1. Eq. 6.71 is Fourier’s law of heat conduction in one dimension. The negative sign indicates that the heat flux is directed from the warmer to the colder end of the layer. The proportionality factor is referred to as the material’s thermal conductivity λ. This is a material constant and independent of the geometry. However, it is dependent on the temperature although in most fluid mechanical calculations, it is often assumed to be a constant. Tab. 6.8 lists a selection of thermal conductivity values for solids. Tab. 9.3 lists a selection of thermal conductivity values for liquids and gases.

Tab. 6.8

Thermal conductivity of selected substances [4, 15]

SubstancesTemperature °CThermal conductivity λ W m−1 K−1Density ρ kg m−3cp J g−1 K−1
diamond251 0003.510.52
copper03908.940.39
aluminum02202.730.90
carbon steel0657.860.49
aluminum oxide (sintered)0353.800.72
aluminum oxide (sapphire)100303.980.76
stainless steel0147.900.49
titanium07.804.500.52
water02.201.004.18
coal200.261.35
pine wood600.260.451.70
graphite400.180.480.72
asbestos1000.100.40
glass00.102.500.67
cork1000.080.05
asphalt200.062.100.92
Polymers
Bakelite201.401.30
celluloid300.021.401.30
polystyrene foam200.030.05
nylon250.301.15
polytetrafluoroethylene250.262.201.00
polyurethane foam200.060.07
shellac200.231.10

t0045

Fourier’s Law of Heat Conduction in Three Dimensions. The temperature profile that is formed due to this difference in temperature must be calculated. It can be found by extending Eq. 6.71 to the infinitesimal volume dV (Fig. 6.5b). We will derive Fourier’s law of heat conduction in three dimensions. The temperature changes along the axes are given by

in x-direction

dTx=Tx+ΔxTx=Tx+TxdxTx=Txdx

si240_e  (Eq. 6.74)

in y-direction

dTy=Ty+ΔxTy=Ty+TydyTy=Tydy

si241_e  (Eq. 6.75)

in z-direction

dTz=Tz+ΔxTz=Tz+TzdzTz=Tzdz

si242_e  (Eq. 6.76)

Substituting Eq. 6.74, Eq. 6.75, and Eq. 6.76 into Eq. 6.71 allows us to rewrite

q˙A,total=(λTxdx+λTydy+λTzdz)=λ(T)

si243_e  (Eq. 6.77)

that is the Fourier law for heat conduction is three dimensions. The operator si244_e is the nabla operator which will be introduced in section 7.1.3.1.

Balance of Heat Flows. We will now turn to the balance of heat flows in the infinitesimal volume (Fig. 6.5c). It is given by

in x-direction

dq˙A,xdydx=q˙x(q˙x+q˙xxdx)dydz=q˙xxdxdydz

si245_e

in y-direction

dq˙A,ydxdy=q˙y(q˙y+q˙yydy)dxdz=q˙yydxdydz

si246_e

in z-direction

dq˙A,zdxdy=q˙z(q˙z+q˙zzdz)dxdy=q˙zzdxdydz

si247_e

in summary

deV,totdxdydz=(q˙xx+q˙yy+q˙zz)dxdydzdeV,tot=(q˙xx+q˙yy+q˙zz)

si248_e  (Eq. 6.78)

where we have used the first law of thermodynamics (see Eq. 6.32) in order to relate the heat flows to the change of energy. Using Eq. 6.71, Eq. 6.78 can be rewritten as

deV,tot=λ(2Tx2+2Ty2+2Tz2)

si249_e  (Eq. 6.79)

=λΔT

si250_e  (Eq. 6.80)

Here we have used the Laplace operator ∆ which will be introduced in section 7.1.3.5.

Stationary Heat Conduction.Stationary heat conduction refers to the fact that we do not take into account warming of the fluid, i.e., there is no change of enthalpy. If the control volume does not have any additional heat sources (e.g., en exothermic chemical reaction) or sink (e.g., a endothermic reaction) the change of total energy Eq. 6.80 becomes a classical Laplace equation (see section 8.1.6)

λΔT=0

si251_e  (Eq. 6.81)

6.9.2.2 Instationary Heat Conduction

Instationary heat conduction takes into account changes of enthalpy of the fluid control volume. We therefore have to use Eq. 6.80 and substitute the left-hand side of the equation by the change of enthalpy over time as a consequence of heat conduction. This equation results in

ρcpTt=λΔTTt=λρcpΔTTt=αΔT

si252_e  (Eq. 6.82)

where we have introduced the thermal diffusivity α which has the unit mm2 s−1. It is composed only of material constants and therefore is also a constant for a given material (see section 9.9.3). Tab. 9.4 lists a selection of thermal diffusivity values for commonly encountered liquids. Eq. 6.82 is the Fourier law for instationary heat conduction in three dimensions.

6.9.3 Diffusion

6.9.3.1 Introduction

Diffusion is the random movement of small particles over time. This effect of the movement of small particles across large distances was first described analytically by German physician Adolf Fick.1 It is one of the two important mass transport phenomena that we encounter in fluid mechanics (the other being convection). Convection is the term used for the bulk movement of liquids, e.g., in the form of flowing fluids. If the fluid is at rest, no mass will be transported. On the other hand, diffusion is the mass transport phenomena that can occur even when the bulk of the liquid is at rest. It is based on the action of Brownian motion which is sufficient to move particles if they are sufficiently small, or light. Therefore, diffusion is only relevant for molecules of sufficiently small size.

6.9.3.2 Visualizing Diffusion: A Digital Diffusion Experiment

Rationale. Many students have difficulty with the concept of diffusion because it seems so different from the Newtonian mechanics we are accustomed to. Usually particles move because they have been exposed to forces, thus they gained momentum which essentially conserved the action of the forces that accelerated them. This type of movement seems intuitive: If we apply a force, a particle will move. Diffusion seems to work on entirely different mechanics which is not entirely true.

Diffusion is actually a very simple process which can be visualized very easily. We will do this using a very simple Maple listing which is shown in listing 6.1. This is in the true sense of the meaning a “digital diffusion” experiment, i.e., a computer program that emulates surprisingly exactly what happens during diffusion.

Listing 6.1

[Maple] Listing for the digital diffusion experiment. A digital version of the listing can be found under the name Diffusion.mw in the supplementary material.

1   restart:read "Core.txt":
2
3   #this function makes the given number of steps using a random number generator
4   dosteps:=proc(generator,start,epsilon,timesteps) local i,traveled;
5    traveled:=0;
6    for i to timesteps do
7     if generator()=1 then
8       #move left
9       traveled:=traveled-epsilon;
10     else
11      #move right
12      traveled:=traveled+epsilon;
13     end if;
14   end do:
15   return start+traveled;
16  end proc:
17
18  #this function creates our diffusion model
19  diffuse:=proc(moleculecount,epsilon,timesteps) local i,generator,output;
20   #create the random number generator
21   generator:=rand(0..1);
22   #this is the output: x-values (first column) and y-values (second column)
23   output:=Matrix(moleculecount,2);
24   for i to moleculecount do
25     #x-value
26     output[i,1]:=dosteps(generator,0,epsilon,timesteps);
27     #y-value
28     output[i,2]:=dosteps(generator,0,epsilon,timesteps);
29   end do:
30   return output:
31  end proc:
32
33  #perform 'digital diffusion'
34  quickplot([diffuse(10,1,10),diffuse(10,1,50),diffuse(10,1,100)],-300..300,-300..300,"x [a.u.]", "y
       [a.u.]",["10 steps","50 steps","100 steps"],true);
35  quickplot([diffuse(10,5,10),diffuse(10,5,50),diffuse(10,5,100)],-300..300,-300..300,"x [a.u.]", "y
       [a.u.]",["10 steps","50 steps","100 steps"],true);
36  quickplot([diffuse(10,10,10),diffuse(10,10,50),diffuse(10,10,100)],-300..300,-300..300,"x [a.u.]",
       "y [a.u.]",["10 steps","50 steps","100 steps"],true);
37  quickplot([diffuse(10,20,10),diffuse(10,20,50),diffuse(10,20,100)],-300..300,-300..300,"x [a.u.]",
       "y [a.u.]",["10 steps","50 steps","100 steps"],true);

Suppose we have a very very small particle, e.g., an ion. If a particle is very small, its mass will also be very small. Even if not accelerated by forces, this particle will eventually start to move due to Brownian motion. Brownian motion is usually neglected because it is very much smaller than Newtonian movement. However, if the particle is very small, the forces associated with Brownian motion can have a certain effect and the particle will start to move. Obviously this motion will increase as the temperature increases. Please note that this movement is very small, especially compared to Newtonian movement. The particle will move around its original position. The extent of this movement will be influenced by a number of factors. The mass and the size of the particle will obviously play a dominant role. Larger and heavier particles will move less compared to lighter and smaller particles. The density and the viscosity of the surrounding also plays a crucial role. In dense and highly viscous liquids, the particle movement is hindered due to higher friction forces. All of these factors can be summarized in a variable we will call img. img is the distance that a particle moves from its original position in a given time interval ∆t. The longer this time interval, the longer the traveled distance img will be.

For the sake of simplicity, we will only consider one-dimensional particle movement, i.e., the particle only travels along one axis (in this case the x-axis). Let us now assume our particle starts at the beginning of our experiment at the origin x (t = 0) = 0 (see Fig. 6.6). Our particle is now free to move either to the left or to the right side. Under perfect conditions, the probability is 0.5 and therefore identical for both sides. Therefore it is a purely random event if the particle will move to x (t = ∆t) = or to x (t = ∆t) = −img. In both cases,

f06-06-9781455731411
Fig. 6.6 Visualization of diffusion as a series of discrete movements of width in the time interval ∆t.

the particle will have moved. Now during the second time interval, the particle is again free to move either to the left or right. Therefore it may end up as x (t = 2∆t) = 2img , x (t = 2∆t) = −2img , or x (t = 2∆t) = 0 where the probability for x = 0 is 0.5, and the probability for x = 2img and x = −2img is 0.25, respectively. The chances are thus higher of finding our particle at the origin after two time steps. Here you can see why diffusion has a probabilistic element to it: The particle’s position is a probability distribution. It is simply more likely to be in the origin than anywhere else but it may also have moved.

Maple Worksheet. If you look at listing 6.1 we can simply emulate this probabilistic movement via a random number generator. This is what the function dosteps does (see line 4). It takes a random number generator generator as arguments as well as a particles original location start. It also requires the information how far a particle may travel. This information is given by the argument epsilon or as we have been using. The last argument timesteps indicates the number of sequential steps we want our particles to move, i.e., the number of time intervals ∆t. The function dosteps iterates the number of time steps. For each time step, it requests a number from the random number generator which will only return 0 or 1. In case of it returning 1, the particle moves to left by decrementing the variable traveled by epsilon. If the random number generator returns 0 the variable traveled is increment by epsilon thus the particle moves toward positive x, i.e., to the right. After the given number of steps, the function returns the sum of the original position start and the traveled distance traveled. Therefore this function effectively emulates a particle which can travel the distance img per time step and returns to its location after timesteps number of timesteps assuming its original position start.

The second function we define is diffuse (see line 19). This function takes three parameters: the number of molecules moleculecount, epsilon, and timesteps the latter two of which we already know. The function first creates a random number generator that only creates the values 0 and 1 with equal distribution. In the next step, it creates a matrix with two columns and a number of lines equivalent to the number of molecules. It then iterates and, for each molecule, it calls the function dosteps both for the x-value (first column) and for the y-value (second column). The molecules are “seeded”, i.e., placed at the origin x = y = 0 at the start of the experiment.

The output of this “digital diffusion experiment” is shown in Fig. 6.7. In this experiment a total number of 20 molecules were seeded at the origin. Depending on the number of timesteps, i.e., the duration of the experiment and the “diffusion step width” img one can clearly see that the cloud of molecules distributes, i.e., smears. This is exactly what happens in diffusion. The plots are created using the code starting from line 19. You will see that every time you run these commands the plots will look slightly differently. This is due to the fact that diffusion is a random process and the output will look slightly different every time.

f06-07-9781455731411
Fig. 6.7 Output of the “digital diffusion experiment”. A total of 20 molecules were seeded at the beginning of the experiment at the origin x = y = 0. Various values for were used and a total of 10, 50, and 100 time steps assumed. The influence of and the time is clearly visible. The higher and the longer the time, the more the molecules distribute over time. This experiment shows the underlying mechanism of diffusion.

Diffusion in More Than One Dimension. It may seem strange that we can simply convert from one-dimensional to two-dimensional diffusion simply by calling the function dosteps for x and y. In fact this is correct because during our assumption of diffusion, we assume the particle to be able to move img in one direction. Obviously we would need to correct the lengths at which the particle can travel in order to ensure that the total distance traveled (which is (Δx)2+(Δy)2si253_e) is equal to img but this does not change the overall picture. We may rewrite the algorithm such that the random number generator yields three values which can be interpreted as

 Indicator if the particle movement in x is in the positive or negative direction (first value)

 Indicator if the particle movement in y is in the positive or negative direction (second value)

 Fraction values that determines at which extent img will be contributed to movement along the x-axis or the y-axis

6.9.3.3 Conservation of Mass

In the following, we will derive the two laws of Fick for diffusion. We start with the first law which describes the mass flux into and out of a control volume. Here, we use the mass flow normalized to a surface which is generally referred to as the mass flux J with the unit kg m−2 s−1. Balancing the mass flux at the unit cell (see Fig. 6.8) we find

f06-08-9781455731411
Fig. 6.8 Derivation of the mass conservation for the first law of Fick by balancing the mass flows at the unit cell.

mt=(Jx(Jx+Jxx))dydz+(Jy(Jy+Jyy))dxdz+(Jz(Jz+Jzz))dxdy=(Jxx+Jyy+Jzz)dxdydz=(Jxx+Jyy+Jzz)dV=JdV

si254_e  (Eq. 6.83)

We divide this equation by dV and obtain the conservation of mass as a function of the mass concentration ρ and the mass flux J

ρt=J

si255_e  (Eq. 6.84)

where we have again used the nabla operator which will be formally introduced in section 7.1.3.1.

Derivation Using the Gauss’s Theorem. We can derive the same equation using Gauss’s theorem (see section 7.2.1) which gives us the amount of substance transported across the boundary of an arbitrary control volume is given by

JndA=JdV

si256_e

which must be equal to the negative change of mass inside of the control volume. Therefore

mt=JdVρt=J

si257_e

6.9.3.4 Fick’s First Law of Diffusion

Diffusion is a process that can be derived from kinetic gas theory. It is simply the relation between the kinetic energy of a particle or a molecule, i.e., its inertia and the friction forces of the surrounding fluid. At very low Reynolds numbers (Re si183_e 1, see section 9.9.8 for an introduction to the Reynolds number) a particle will experience a friction proportional to its velocity

v=αF

si258_e  (Eq. 6.85)

where σ is referred to as mobility. The mobility is inverse-proportional to the size of the molecule, i.e., its diameter d and the viscosity η of the fluid: The higher the diameter and the more viscous the fluid, the stronger the mobility will be hindered. The drag that the surrounding fluid will have on the particle is given by Stoke’s1drag law which states that

Fdrag,sphere=6πηrσ=16πηr

si259_e  (Eq. 6.86)

In general, a chemical potential (which has the unit of energy) is dependent on the local concentration of a compound according to

μ=μ0+kBTlncc0

si260_e

If this potential has a gradient, it will create a force according to

F=μ=kBTcc

si261_e  (Eq. 6.87)

Using Eq. 6.85, Eq. 6.86, and Eq. 6.87 we find

v=σF=16πηrF=kBT6πηrccvc=kBT6πηrcJ=kBT6πηrcJ=Dc

si262_e  (Eq. 6.88)

where we have introduced the massflux J with the unit kg m−2 s−1 and the diffusion or diffusivity coefficient D as

D=kBT6πηr

si263_e

The mass flux is the amount of substance transported through a given surface area per time. D gives a relation of how well a gradient in concentration can be “turned into” mass movement. For low mobility molecules, e.g., heavy or bulk molecules, as well as for highly viscous environments, these values are small. Eq. 6.88 is referred to as Fick’s first law of diffusion. It states that the mass flux is proportional to the gradient in concentration.

Returning to our experiment with the “digital diffusion”, this is surprisingly straightforward. We assume the particle movement to be equally probable in the positive and the negative x-direction. However, assume there is a larger conglomeration of particles in the negative x-direction. As the volume is more “crowded” in this direction, the particle movement in the other direction is more probable. The movement of the particle is hindered by the higher density of particles in the negative x-direction. An everyday analogy to this is a traffic jam: The more cars, the slower the movement and in consequence, fewer cars pass by per unit time. Obviously, the direction of the particle movement is in the opposite direction of the concentration gradient, i.e., away from it (thus the negative sign).

The proportionality factor for this relation is generally referred to as the diffusion constant D. It has the unit m2 s−1 which is chosen such that the units of the equation match up. Referring to our “digital diffusion” experiment, it may be interpreted as a combination of the time step ∆t and the stepwidth img. The higher this value, the faster a molecule can move. Obviously, mass transport will be significantly faster at higher diffusion constants.

Diffusion Time. Mass transport by diffusion is not a fast process. The root-mean-square position of a particle after a given time interval t, i.e., the distance the particle has traveled in this time interval is given by

x¯=2tD

si264_e  (Eq. 6.89)

The mathematics of this relation is derived in section 8.3.8.5. In general the distance traveled by molecule scales with the square root of both the time and the diffusion coefficient. Tab. 6.9 lists some commonly encountered diffusion coefficients. Fig. 6.9 displays the difference in mobility of molecules of different size. As you can see, mass transport by diffusion becomes a very slow process for bulkier molecules.

f06-09-9781455731411
Fig. 6.9 Diffusion times required for traveling the length x calculated for different molecules in water. Diffusion values: Dethanol = 1.24 × 10−5 cm2 s−1, DDNA,30 bp = 4.23 × 10−6 cm2 s−1, DDNA,10 000 bp = 6.4510 × 10−8 cm2 s−1.

Tab. 6.9

A selection of diffusion constants of substances in water

SubstanceDiffusion coefficient D 10×10-5 cm2 s−1Temperature °CSource
methanol1.2825[4]
toluene0.8525[4]
ethanol1.2425[4]
glycerol1.0625[4]
deoxyribonucleic acid (DNA)4.90×10-6 × bp−0.7237[17]

t0050

6.9.3.5 Fick’s Second Law of Diffusion

Fick’s second law of diffusion combines the mass conservation equation with the first law of Fick (see Eq. 6.84 and Eq. 6.88). It yields the relationship between the gradient in the mass concentration and the change of mass as

ρt=D2ρ=DΔp

si265_e  (Eq. 6.90)

Again, the Laplace operator ∆ will be formally introduce in section 7.1.3.5. Eq. 6.90 is a homogeneous second-order PDE which is simply referred to as the diffusion equation. In section 8.3.8 we look into methods of solving this equation.

6.9.3.6 Combined Convection and Diffusion

In many practical fluid mechanical problems, diffusion is often neglected due to the slow mass transport over long length scales. Mass transport is therefore solely based on the in- and outflow of mass into and out of the control volumes. This fluid movement is generally referred to as convection. However, in many microfluidic applications, the length scales are reasonably small and it may be necessary to account for both diffusion and convection. We have already derived Fick’s second law of diffusion (see Eq. 6.90) which is the most general PDE for describing diffusion.

6.10 SUMMARY

In this section, we have introduced some of the most important concepts from thermodynamics, chemistry, as well as from heat and mass transfer which we will require for deriving the fundamental equations of fluid mechanics. Traditionally, these are topics with which many students struggle. However, as we have seen, the underlying concepts are not very complicated and the fundamentals can be readily applied once some of the basics have been understood.

1.06

References

[1] Dalton J. A new system of chemical philosophy. 1808 London, (cit. on p. 93).

[2] Laurent L.A. Traité élémentaire de chimie. 1789 (cit. on p. 93).

[3] Rathore M.M. Thermal engineering. Tata McGraw-Hill Education; 2010 (cit. on p. 101).

[4] Lide D.R. CRC handbook of chemistry and physics. 84th ed. CRC press; 2003 (cit. on pp. 102, 107, 115, 124, 126, 134).

[5] Maxwell J.C. A Dynamical Theory of the Electromagnetic Field. Proceedings of the Royal Society of London. 1863;13:531–536 (cit. on p. 103).

[6] Boltzmann L. Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen. Kinetic Theory. 1872;2:88–174 (cit. on p. 104).

[7] Hess H. Thermochemische Untersuchungen. Annalen der Physik. 1840;126(6):385–404 (cit. on p. 111).

[8] Riedel E., Janiak C. Anorganische Chemie. Walter de Gruyter; 2011 (cit. on pp. 112, 113).

[9] Carnot L., Deterville J.-F.-P., Crapelet C., Cloquet J.B.A. Principes fondamentaux de l’équilibre et du mouvement. Paris: Deterville; 1803 (cit. on p. 112).

[10] Marsh K.N., Marsh K. Recommended reference materials for the realization of physicochemical properties. UK: Blackwell Scientific Publications Oxford; 1987 (cit. on p. 115).

[11] van’t Hoff M.J.H. Etudes de dynamique chimique. Recueil des Travaux Chimiques des Pays-Bas. 1884;3(10):333–336 (cit. on p. 124).

[12] Knuuttila T., Tuoriniemi J., Lefmann K., Juntunen K., Rasmussen F., Nummila K. Polarized Nuclei in Normal and Superconducting Rhodium. Journal of Low Temperature Physics. 2001;123(1-2):65–102 (cit. on p. 124).

[13] Sahai R., Nyman L.-Å. The Boomerang Nebula: The Coldest Region of the Universe? The Astrophysical Journal Letters. 1997;487(2):L155 (cit. on p. 124).

[14] Narasimhan T.N. Fourier’s heat conduction equation: History, influence, and connections. Reviews of Geophysics. 1999;37(1):151–172 (cit. on p. 125).

[15] Engineering Toolbox. 2015. (cit. on p. 126) http://www.engineeringtoolbox.com.

[16] Fick A. Ueber diffusion. Annalen der Physik. 1855;170(1):59–86 (cit. on p. 128).

[17] Lukacs G.L., Haggie P., Seksek O., Lechardeur D., Freedman N., Verkman A. Size-dependent DNA mobility in cytoplasm and nucleus. Journal of Biological Chemistry. 2000;275(3):1625–1629 (cit. on p. 134).


1 John Dalton was an English scientist who is generally considered to be the originator of the modern atomic theory. He suggested this “new chemical philosophy” in 1808 [1].

1 Amedeo Avogadro was an Italian scientist who first suggested that the volume (and the mass) of a gas should be proportional to the number of atoms or molecules, irrespective of the type of gas.

1 James Clerk Maxwell was a Scottish physicist who formulated the fundamental electromagnetic equations commonly referred to as Maxwell equations[5]. Maxwell is generally considered as one of the most important scientists of all time.

1 Ludwig Boltzmann was an Austrian physicist who studied the movement of gases. Among others, he derived the characteristic Boltzmann distribution in 1872 [6].

1 Henri Hess was a Russian chemist who made important contributions to the thermodynamic fundamentals of modern chemistry. Among others, he described the law generally referred to as Hess’s law in his seminal work dating from 1840 [7].

1 Lazare Carnot was a French mathematician who is most commonly known for his assumption made in 1803 that any transfer of useful energy will result in energy dissipation due to collision of the particles within the system and the inherent loss of momentum [9]. Therefore any process will dissipate energy as a consequence of inner friction. Entropy is therefore defined as the amount of energy that cannot be used to do useful work. From this he derived the concept of entropy which gives rise to the second law of thermodynamics.

1 Frederick Trouton was an Irish chemist. He developed Trouton’s rule which allows estimating the entropy changes of liquids during phase transition.

1 Hermann von Helmholtz was a German mathematician and physicist. He originally studied medicine at the Friedrich-Wilhelm-Institut in Berlin as his father considered physics to be a field which could not support a living. While being trained for medical service, he studied math and physics and made important contributions in the field of hearing and seeing, effectively blending medicine and the natural sciences. He invented numerous instruments in optics and made important contributions in the field of microscopy. He is considered the academic founding father of modern meteorology. He also contributed significant concepts in electrodynamics, inventing the Helmholtz coil, a specific coil setup which allows very homogeneous electrical fields. The German Helmholtz Association of German Research Centers is named after him.

1 Jacobus van’t Hoff was a Dutch chemist who formulated the Van’t Hoff equation in 1884 [11]. This equation allows estimating the change of reaction speed as a function of the temperature.

1 Adolf Fick was a German physician who is most known for his work on diffusion of substance across membranes which he published in 1855 [16]. The fundamental laws of diffusion are named after him.

1 George Gabriel Stokes was an English physicist who made important contributions to fluid mechanics. Among others, he derived the Stokes’ friction approach which is an important simplification in the derivation of the Navier-Stokes equation. The latter equation carries his name because Stokes made important improvements to the equation originally described by Navier, and Stokes provided a more sound continuum mechanical basis which renders the equation easier to solve.

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

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