99
 p. 1 of 5 At 27°C the reading on a manometer filled with mercury is 60.5 cm. The local acceleration of gravity is 9.784 m s -2 . To what press ure does this height of mercury correspond? Solution: The pressure exerted by the column of mer cury is equal to its height times its density times the local acceleration of gravity, or in symbols P  = h  ρ  g . We have h = 60.5 cm = 0.605 m. The density of mercury at 27°C is 13.53 g cm -3  = 13530 kg m -3  (where would I find this?). So, we have  P  = 0.605 m × 13530 kg m -3  × 9.784 m s -2  = 80088 kg m -1  s -2  = 8.01×10 4  Pa = 0.801 bar Work: Usually, mechanical work can be defined as the product of a force times a distance or the integral of a force over some distance through which it acts. If F  is the component of force along the direction of motion and dl is a differential displacement in that direction, then this can be expressed as dW  = F  dl  LECTURE NOTES CHEMICAL ENGINEERING THERMODYNAMICS What is thermodynamics? One answer: a field of science and en gineering that describes, macrosco pically, the driving forces for the flow of energy (as heat and/or work) and the conditions for equilibrium within a system (where equilibrium is the state where these driving forces are absent and the system has no t endency to change). Thermodynamics does not deal with rates of change, but only with the driving forces that cause change. To do thermodynamics, we must understand basic physics (like that learned in the mechanics portion of a first-year college physics course) and be adept at using and converting un its of measurement . In engineering , most numb ers are meaningless without uni ts. We will not review units and un it conversions here, but y ou should read the review of dimensions and units in chapter 1 of Smith, Van Ness, and Abbott (SVA). Pay particular attention t o the section on pressure. In general, while I will post detailed lec ture notes like the ones you are reading, and these will follow the textbook fairly closely, these are not an adequate substitute for also reading the textbook chapters as we cover them. Example 1.3 from SVA

Thermodynamics Lecture Notes

Embed Size (px)

Citation preview

  • p. 1 of 5

    At 27C the reading on a manometer filled with mercury is 60.5 cm. The local acceleration of gravity is 9.784 m s-2. To what pressure does this height of mercury correspond? Solution: The pressure exerted by the column of mercury is equal to its height times its density times the local acceleration of gravity, or in symbols P = hg. We have h = 60.5 cm = 0.605 m. The density of mercury at 27C is 13.53 g cm-3 = 13530 kg m-3 (where would I find this?). So, we have P = 0.605 m 13530 kg m-3 9.784 m s-2 = 80088 kg m-1 s-2 = 8.01104 Pa = 0.801 bar Work: Usually, mechanical work can be defined as the product of a force times a distance or the integral of a force over some distance through which it acts. If F is the component of force along the direction of motion and dl is a differential displacement in that direction, then this can be expressed as dW = F dl

    hclTypewritten textLECTURE NOTES

    hclTypewritten TextCHEMICAL ENGINEERING THERMODYNAMICS

    hclTypewritten Text

    hclTypewritten Text

    hclTypewritten Text

    hclTypewritten TextWhat is thermodynamics? One answer: a field of science and engineering that describes, macroscopically, the driving forces for the flow of energy (as heat and/or work) and the conditions for equilibrium within a system (where equilibrium is the state where these driving forces are absent and the system has no tendency to change). Thermodynamics does not deal with rates of change, but only with the driving forces that cause change. To do thermodynamics, we must understand basic physics (like that learned in the mechanics portion of a first-year college physics course) and be adept at using and converting units of measurement. In engineering, most numbers are meaningless without units. We will not review units and unit conversions here, but you should read the review of dimensions and units in chapter 1 of Smith, Van Ness, and Abbott (SVA). Pay particular attention to the section on pressure. In general, while I will post detailed lecture notes like the ones you are reading, and these will follow the textbook fairly closely, these are not an adequate substitute for also reading the textbook chapters as we cover them. Example 1.3 from SVA

    hclTypewritten Text

    hclTypewritten Text

    hclTypewritten Text

    hclTypewritten Text

    hclTypewritten Text

    hclTypewritten Text

  • where dW is the differential amount of work performed when the force F acts through the displacement dl.

    In thermodynamics, we often deal with the work done by an expanding fluid (or work done on a fluid to compress it). In this case, the force F is the pressure multiplied by the area over which the pressure is applied (F = PA). If the compression is done by a cylinder of constant area A then the change in total volume of the fluid is dVt = A dl. Using these expressions for dl and F gives

    dW = -P dV t

    The cross-sectional area cancels out (so the work is independent of the shape of the piston and cylinder). The negative sign arises if we use the sign convention that work done on the gas by the piston is positive. When the piston does work on the gas, the change in volume is negative and the pressure is positive, so the negative sign is required to make the work come out to be positive. In integral form, then, the work on a gas in compressing it from pressure P1 to pressure P2 is:

    2

    1

    P

    PW PdV=

    How would we compute this, though? The limits of the integral are in terms of pressure, while the variable of integration is dV. We need an equation of state that relates the pressure and volume (and temperature, if temperature changes) in order to evaluate this. We will spend substantial time this semester considering equations of state.

    Forms of Mechanical Energy (Kinetic Energy and Gravitational Potential Energy)

    Kinetic energy (EK = mu2 where m is the mass of an object and u is its speed) and gravitational potential energy (EP = mgz where m is the mass of an object, z is its elevation relative to a reference elevation, and g is the local acceleration of gravity) are forms of mechanical energy that can be converted (fully) to work. Note that only changes in or relative amounts of kinetic and potential energy are meaningful, since the objects speed and elevation must be defined relative to some stationary reference frame and reference position, respectively.

    Mechanical Energy Conservation: Example 1.4 in SVA.

    An elevator with a mass of 2,500 kg rests at a level 10 m above the base of an elevator shaft. It is raised 100m above the base of the shaft, where the cable holding it breaks. The elevator falls freely to the base of the shaft and strikes a strong spring. The spring is designed to bring the elevator to rest and, by means of a catch arrangement, to hold the elevator at the position of maximum spring compression. Assuming the entire process to be frictionless, and taking g = 9.8 m s-2, calculate:

    (a) The potential energy of the elevator in its initial position, relative to the base of the shaft.

    (b) The work done in raising the elevator

    Lecture 1

  • (c) The potential energy of the elevator in its highest position relative to the base of the shaft.

    (d) The velocity and kinetic energy of the elevator just before it strikes the shaft.

    (e) The potential energy of the compressed spring.

    (f) The energy of the system consisting of the elevator and spring (1) at the start of the process, (2) when the elevator reaches its maximum height, (3) just before the elevator strikes the spring, and (4) after the elevator has come to rest.

    Solution:

    The whole situation will probably be clearer if we draw a sketch of it:

    10 m

    100 m

    (a) If our reference elevation is the bottom of the shaft, and the initial elevation of the elevator is 10 m above the bottom of the shaft, then the potential energy of the elevator (relative to the bottom of the shaft) is

    Ep = mgz = 2500 kg 9.8 m s-2 10 m = 245,000 kg m2 s-2 = 245 kJ

    (b) The work done to raise the elevator is equal to the integral of the force applied to the elevator over the distance it is raised. In this case, the force is constant F = mg (or more formally, F = mg/gc to make it work in English units). So, the work done is

    ( )2 2

    1 1

    2 1

    z z

    z z

    W Fdz mgdz mg z z= = = W = 2500 kg 9.8 m s-2 90 m = 2,205,000 kg m2 s-2 = 2,205 kJ

    Lecture 1

  • (c) Just as in part (a), we have

    Ep = mgz = 2500 kg 9.8 m s-2 100 m = 2,450,000 kg m2 s-2 = 2,450 kJ

    (d) We use the principle of conservation of mechanical energy (in this frictionless system) to equate the change in gravitational potential energy to the change in kinetic energy.

    ( ) ( ),2 -2

    ,

    0

    0 2,450 kJ 0 0

    2,450 kJ = 2,450,000 kg m s

    P K

    K after

    K after

    E E

    E

    E

    + =

    + =

    =

    That is, the 2,450 kJ of potential energy that the elevator has when it is at a height of 100 m and its kinetic energy is zero (because its velocity is zero) is converted completely to kinetic energy when it falls to a height of zero (where its potential energy is zero). The velocity corresponding to this kinetic energy is obtained from

    ( )

    2 -2 21, 2

    2 -2

    12

    2,450,000 kg m s

    2,450,000 kg m s 44.3 m/s2500

    K afterE mu

    ukg

    = =

    = =

    Note that we could get this same result by solving Newton's laws of motion for the elevator. At t = 0, we have z = 100 m and u = 0. The acceleration of the elevator (by gravity) is a = g = 9.8 m s-2. Integrating this to get the velocity gives:

    -29.8 m s

    , where C is a constant of integration

    dv adtv at C

    = =

    = +

    At t = 0, v = 0, so C = 0, and v = -9.8 t m/s

    Integrating again,

    -2

    2 212

    9.8 m s

    4.9 , where C is a constant of integration

    dz v tdtz vt C C t

    = =

    = + =

    At t = 0, z = 100 m, so C = 100 m, and z = (100 - 4.9t2) m

    To find the velocity when the elevator reaches z = 0, we first must find the time when z = 0 by solving 100 - 4.9t2 = 0, which gives t = 4.518 s. At this time, the velocity is then v = -9.8 m s-2 4.517 s = 44.3 m/s. We could show (but won't) that the law of conservation of mechanical and potential energy can be derived from Newton's laws of motion.

    This is one of many examples in the course in which there is more than one logical and appropriate way to reach the correct answer. On occasion, I will show methods that differ from

  • those in the textbook. In such cases, you are encouraged to try applying both methods, which will hopefully deepen your understanding of the problem and its solution.

    (e) The change in potential energy of the spring plus the change of kinetic energy of the elevator must sum to zero (again, for a frictionless system), so EP,spring = 2,450 kJ after the elevator comes to rest (converting all of its kinetic energy to potential energy of the spring). Note that since the spring is described as 'very stiff', we are neglecting any change in the elevation of the elevator during the compression of the spring. If the spring compresses a significant distance, then the gravitational potential energy would become negative (since the elevation of the elevator would be less than the reference elevation) and the potential energy of the spring would be slightly more than 2,450 kJ.

    (f) If the elevator and spring are taken together as the system, then the energy of the system only changes when something outside the system does work on it, which only happens when the elevator is initially raised from 10 m to 100 m. So, initially, the total energy is 245 kJ (as in part (a)). After the elevator is raised to 100 m, the total energy of the spring plus the elevator is 2,450 kJ. When the elevator falls and compresses the spring, energy is converted from one form to another within the system, but no work is done on or by objects outside the system, so the total energy remains 2,450 kJ.

    Heat: See discussion in SVA section 1.9.

    Heat is what flows when energy is transferred from a warmer object to a cooler one, raising the internal energy of the cooler object and lowering the internal energy of the warmer one, until their temperatures are equal. A more precise definition will require improved definitions of temperature and internal energy. The first law of thermodynamics (up next) will show us that heat and work are equivalent in some senses but not in others. As a result, heat is measured in the same units (joules or btu) as work.

  • When we add heat to or do work on a substance, we may increase its internal energy. From a macroscopic point of view, this is energy stored within the material that can later be recovered as heat transferred to a cooler object. Microscopically, this energy is stored in the motions of the molecules that make up the material and in the potential energy of the interactions between the molecules. In classical thermodynamics, absolute values of internal energy aren't known, just relative values, which are all that are required for thermodynamic analysis anyway.

    The First Law: Energy is Conserved

    Although energy assumes many forms, the total quantity of energy is constant, and when energy disappears in one form it appears simultaneously in other forms.

    We often analyze problems by dividing the universe into the system (the part we are explicitly studying) and the surroundings (everything else).

    Any changes in the energy of the system must be accompanied by changes in the energy of the surroundings. That is

    (Energy of the System) + (Energy of the Surroundings) = 0

    Energy Balances for Closed Systems:

    If our system is closed (no mass flows across its boundaries) then energy can only enter or leave the system as heat or as work, and we have

    (Energy of the System) = Q + W

    The System (the part of the universe

    were looking at)

    The Surroundings (Everything Else)

    Q

    W

    FIRST LAW OF THERMODYNAMICS (chapter 2 in SVA)

    Internal energy:

  • Where Q is energy transferred from the surroundings to the system by heat, and W is energy transferred from the surroundings to the system as work (work done on the system by the surroundings). This implies that

    (Energy of the Surroundings) = -Q - W

    The sign convention for Q and W used here is the one used in SVA, but is not universal. You should always be sure you understand which way energy is flowing and not just rely on the sign in the equations to get it right.

    If all energy transferred into the system goes to increase the internal energy of the system (as opposed to increasing its kinetic or potential energy by changing its overall position or velocity), then we have

    Ut = Q + W

    where Ut is the total internal energy of the system.

    For infinitesimally small (differential) changes, this is

    dUt = dQ + dW

    The quantity Ut is an extensive quantity - it is the total internal energy of the system, which if the system is homogeneous, is proportional to the total amount of matter in the system.

    Homogeneous means "the same throughout" - at every point within the system the properties (temperature, pressure, density, composition, etc.) are the same as at every other point.

    If the system is homogeneous, then we can define intensive properties like the internal energy per mole or per mass

    U = Ut/n = internal energy per mole = total internal energy divided by the number of moles

    U = Ut/m = internal energy per mass = total internal energy divided by the total mass

    Then, for a homogeneous closed system containing n moles of material, we can write

    (nU) = nU = Q + W

    d(nU) = ndU = dQ + dW

    Often, we simply write thermodynamic equations for a representative amount of a homogeneous material, in which case, we would just write (for one mole of material, for example)

    U = Q + W

    dU = dQ + dW

    hclTypewritten Text

  • This last expression, in particular, allows us to keep track of the molar internal energy of a substance and compute changes in it by computing the amount of work done on it by its surroundings and the amount of heat that flows to it from its surroundings. By observing how its temperature and other properties change as its internal energy changes we can also relate changes in internal energy (a property of the substance) to changes in its other properties (like density, etc.) that are observable.

    Thermodynamic State and State Properties

    At a given thermodynamic state, a homogenous system always has the same thermodynamic properties. For a homogeneous pure substance, fixing two properties (i.e. temperature and pressure or temperature and density or density and pressure) fixes all of its other thermodynamic properties (temperature, pressure, density, internal energy, etc.). For more complex systems, more than 2 thermodynamic properties may have to be specified to specify the state of the system, but once the required number is specified, the thermodynamic state of the system is fixed and all other thermodynamic properties are also fixed.

    State functions depend only on the thermodynamic state of the system and do not depend on the path by which the system arrived at that state. Internal energy is a state function.

    Q and W are not state functions. They do not represent properties of the system, but represent interactions of the system with its surroundings. As such, they depend on the path. The system can get from one state to another through different paths, where by different paths we mean through different interactions with its surroundings. If two paths move the system between the same two states (they cause the same change in the state of the system) then the sum Q + W will be the same for both paths, since the change in internal energy must be the same. However, Q and W, individually, will be different for the two paths.

    For example, suppose I have some water at a temperature of 280 K and a pressure of 1 atmosphere. Fixing the temperature and pressure specifies the state of the system. I could stir the water violently until its temperature increased to 300 K (still at 1 atmosphere) while keeping it thermally insulated. In this case, Q would be zero, and we would have U = W. Alternatively, I could heat the water from 280 K to 300K by sitting the (non-insulated) beaker on a warm surface. In this case, no work would be done on the water (W = 0), and we would have U = Q. Both methods of heating the water result in the same change in the state of the water (they go from the same initial thermodynamic state to the same final thermodynamic state), so U is the same in both cases. However, Q and W are different for the two cases.

    The Phase Rule:

    For any non-reacting system at thermodynamic equilibrium, the number of independent variables that must be fixed to establish its intensive state is given by

    F = 2 - + N

    where is the number of phases present, N is the number of chemical species present, and F is the number of degrees of freedom that the system has. The number of degrees of freedom of the

  • system is the number of phase rule variables (temperature, pressure, composition of each phase, etc.) that must be fixed to completely specify the state of the system, including all other phase-rule variables.

    See example 2.5 in SVA.

    Reversible Processes

    In thermodynamic analyses, we often make use of the idealized concept of a reversible process, much like in first-year physics we make use of a lot of frictionless processes, even though such a process can only be approximately realized in a real system.

    A reversible process is one whose direction can be changed at any point by infinitesimal changes in external conditions

    (which means there must be no friction, no inertia (so it must happen infinitely slowly), no heat transfer through finite temperature differences, and so on).

    A reversible process

    Is frictionless

    Is never more than infinitesimally away from equilibrium

    Goes through a set of equilibrium states

    Is driven by infinitesimally small force or temperature imbalances

    Can be reversed at any point by an infinitesimal change in external conditions

    When reversed, retraces its original path back to the original state of the system

    Constant Volume processes, Constant Pressure processes, and Enthalpy

    For a homogeneous, closed system containing n moles of a substance, the energy balance was (in differential form)

    d(nU) = dQ + dW

    For a mechanically-reversible closed-system process, the work term is

    dW = -PdVt = -Pd(nV) = -nPdV

    So, for a constant volume process, the work is zero and we have

  • dQ = n dU (at constant V)

    or, in integral form

    Q = nU (at constant V)

    At constant pressure, the energy balance becomes

    d(nU) = dQ - nPdV (at constant pressure)

    or

    dQ = n(dU + PdV) = nd(U + PV) (at constant pressure)

    This motivates the definition of enthalpy as H = U + PV, so that at constant pressure dQ = n(dH), just as at constant volume dQ = n(dU).

    Note that even though Q is not a state function, H is defined in terms of U, P, and V, which are state functions, so H is a state function as well.

    See example 2.8 in SVA.

    Heat Capacity

    The constant volume heat capacity is defined as

    vV

    dUCdT

    For a constant volume process, we can then write

    dU = CvdT (at constant V)

    Or, integrating over a finite change in temperature

    2

    1

    T

    v

    T

    U C dT = (at constant V) For a mechanically reversible constant-volume process, we then have

    2

    1

    T

    v

    T

    Q n U n C dT= = (for a mechanically reversible, constant volume process) Likewise, we define the heat capacity at constant pressure as

  • pP

    dHCdT

    Then, for a constant pressure process on a fixed amount of a homogeneous substance, we have

    dH = CpdT (at constant P)

    and for finite temperature changes

    2

    1

    T

    p

    T

    H C dT = (at constant P) And finally, for a mechanically reversible, constant pressure process, this tells us that

    2

    1

    T

    p

    T

    Q n H n C dT= = (for a mechanically reversible, constant pressure process) See example 2.9 in SVA

    Mass and Energy Balances for Open Systems

    An open system is one in which mass, as well as energy, crosses the system boundaries. For an open system, energy is carried in and out of the system with the mass that enters and leaves the system, as well. In general, we can write a mass balance on a system as

    accumulation = in - out

    Short of nuclear reactions, we cannot have a 'generation' term in a mass balance (mass is conserved). The same is true for a mole balance in a non-reacting system, but of course in a reacting system we can have a change in the total number of moles due to reaction (leading to a 'generation' term in the balance equation).

    The total mass balance on a fixed control volume with streams entering and leaving it can be written in SVA's notation as

    ( ) 0cv fsdm mdt + =

    Where the first term is the time derivative of the total mass in the control volume (accumulation) and the second term is the difference between the total inflow and outflow (out - in) in the various streams. This should be very simple compared to things done last semester in CE 212, so don't let it seem harder than it really is!

  • Similarly, the energy balance can be written as

    ( ) ( )( )212cv fsd mU H u zg m Q Wdt + + + = +

    where U is the internal energy per unit mass (specific internal energy) within the (homogeneous) control volume. H is the specific enthalpy of each inflow and outflow stream, u is the average flow velocity of each inlet and outlet stream, and z is the elevation of each inlet and outlet stream. In many, many cases, kinetic energy and gravitational potential energy can be neglected, and this becomes simply

    ( ) ( )cv fs

    d mUmH Q W

    dt+ = + (when kinetic and potential energy changes are negligible)

    Furthermore, we are often interested in systems at steady state, in which case the time derivative is zero, and we have simply

    ( ) fsmH Q W = + (at steady-state AND when kinetic and potential energy changes are negligible)

    There are many possible variations on this energy balance, and one must carefully assess, in a given problem which form should be used (which terms should be included).

    Example 2.12 in SVA

    Example 2.15 in SVA

    hclTypewritten textUnsolved Problems (SVA)prob 2.6 to 2.15 and 2.25 to 2.27

  • You should be able to define, and locate on such a diagram, the solid region, liquid region, gas region, supercritical fluid region, fusion curve (melting curve), sublimation curve, vaporization curve, critical point, and triple point.

    Here is the PT diagram for water, which has the unusual feature of a fusion curve with a negative slope:

    PVT behavior of pure substances

    Let's start by looking at PT and PV diagrams for a 'typical' substance. Here is a PT diagram:

    Volumetric Properties of Pure Fluids (chapter 3 in SVA)

  • The PT diagrams don't contain any information about the volume of the fluid, which is an important piece of information for many thermodynamic analyses. So, we should look at some PV diagrams too. Here is a generic one that I drew myself. Again, you should be able to identify all of the regions

    An isotherm (path of constant temperature) on such a diagram might look like this:

    The temperature for that isotherm is far below the critical temperature, so as the pressure is decreased, there is first a small increase in the specific volume (as the liquid phase expands just a

    Solid + Vapor

    Sol

    id

    Sol

    id +

    Liq

    uid

    Liquid + Vapor

    Liquid

    Vapor

    Specific Volume

    Pre

    ssur

    e

    critical point

    Solid + Vapor

    Sol

    id

    Sol

    id +

    Liq

    uid

    Liquid + Vapor

    Liquid

    Vapor

    Specific Volume

    Pre

    ssur

    e

    critical point

  • little), then there is a big change in specific volume at constant pressure, corresponding to the phase change from liquid to gas (note that this will require substantial heat input to keep the temperature constant). After that, the specific volume continues to change with pressure as the gas expands.

    For a temperature above the critical point, the isotherm would look more like

    At temperatures above the critical point, there is no vapor-liquid phase transition. An isotherm just at the critical temperature would have a point of inflection at the critical point, as shown here:

    Solid + Vapor

    Sol

    id

    Sol

    id +

    Liq

    uid

    Liquid + Vapor

    Liquid

    Vapor

    Specific Volume

    Pre

    ssur

    e

    critical point

    Solid + Vapor

    Sol

    id

    Sol

    id +

    Liq

    uid

    Liquid + Vapor

    Liquid

    Vapor

    Specific Volume

    Pre

    ssur

    e

    critical point

  • Equations of State:

    For a single-component, single-phase substance, we know from the phase rule (and from the diagrams at which we were just looking) that two intensive thermodynamic quantities can be specified, and then the state of the system (and therefore all other intensive thermodynamic quantities) is specified. This implies that for a single-phase, single-component system there exists a unique relationship between the temperature, pressure, and specific volume (T, P, and V). We call this relationship an equation of state for the substance. Once we specify two of the three values of P, V, and T, the third can be found from the equation of state. We can consider any one of the three quantities to be a function of the other two. For example, we can consider specific volume to be a function of temperature and pressure

    V = V(T,P), which implies that (by definition of the partial derivatives)

    P T

    V VdV dT dPT P

    = +

    The partial derivatives are related to the isothermal compressibility and the volume expansivity, which are physical properties that are sometimes available from tables in handbooks (for liquids and solids). Volume expansivity () is defined as

    1P

    VV T

    =

    Isothermal compressibility () is defined as

    1T

    VV P

    =

    So, we can write

    dV dT dPV

    =

    If the isothermal compressibility and the volume expansivity were both zero, then the fluid would be incompressible, which is a common approximation for liquids. Using constant values for and is the next-better approximation, which accounts, on average, for the pressure and temperature dependence of the specific volume of liquids. An exact treatment would require pressure and temperature dependent values of and (but this approach is rarely required). See example 3.1 in SVA

    For gases, we are familiar with the ideal gas equation of state, which can be written as

    RTVP

    =

  • or

    RTPV

    =

    or

    PVTR

    =

    Next, we'll consider more realistic (and therefore mathematically more complex) equations of state for gases.

  • At sufficiently low pressure, all gases approach ideal gas behavior, which means that they obey the ideal gas equation of state:

    PV = RT

    where V is the molar volume (volume per mole). The ideal gas equation of state provides the basis for virtually all equations of state for gases. Equations of state that are more accurate at higher pressures usually add corrections to the ideal gas equation of state. Since it is the baseline against which other equations are derived, we will start by going through important properties of ideal gases and and how to do process calculations on them.

    An important property of an ideal gas is that its internal energy depends only on temperature (and not on pressure). This, in turn, implies that the constant volume heat capacity is a function only of temperature (and not of pressure or specific volume). That is

    ( )

    ( )v vV

    U U TU dUC C TT dT

    =

    = =

    From the definition of enthalpy and the constant pressure heat capacity, we can show that H is also a function only of temperature, and we can derive the relationship Cp = Cv + R as follows:

    ( ) ( )

    ( )p vP

    H U PV U T RT H TH dH dUC R C T RT dT dT

    + = + =

    = = + = +

    Because enthalpy and internal energy are functions only of temperature, we have

    dU = CvdT

    and dH = CpdT

    for all processes (not just the usual constant volume and constant pressure processes, respectively).

    So, for an ideal gas undergoing any process we have

    2

    1

    2

    1

    T

    v

    T

    T

    p

    T

    U C dT

    H C dT

    =

    =

    Ideal Gas Equation,

  • regardless of the details of the process.

    Process calculations with ideal gases

    Calculating the heat and work required to affect changes in an ideal gas is particularly simple (at least compared to similar calculations with more complex equations of state). For an ideal gas in a closed system undergoing a mechanically reversible process, we have (on a per mole or per mass basis)

    dU = CvdT = dQ + dW

    We also have (again for a mechanically reversible, closed-system process)

    dW = -PdV

    So we can write

    dQ = CvdT + PdV

    We can rearrange these relationships and use the ideal gas law to eliminate one of the 3 variables (P, V, or T) to solve particular problems. If we substitute P = RT/V, we get

    dW = -(RT/V) dV

    dQ = CvdT + (RT/V) dV

    If we substitute V = RT/P, from which 2R RTdV dT dPP P

    = , we get

    RTdW RdT dPP

    = +

    v pRT RTdQ C dT RdT dP C dT dPP P

    = + =

    Finally, if we substitute T = PV/R, from which V PdT dP dVR R

    = + , we get

    dW = -PdV

    pvvC PC VV PdQ C dP dV PdV dP dV

    R R R R

    = + + = +

    We can use these relationships to solve for the heat and work requirements for all sorts of processes.

    Isothermal Processes:

  • Since the internal energy and enthalpy of an ideal gas are functions only of temperature, they do not change in an isothermal process

    U = H = 0 for an isothermal process

    Using the equations from above with temperature and volume as the independent variables:

    dW = -(RT/V) dV

    dQ = CvdT + (RT/V) dV = (RT/V) dV

    We can integrate to get the following expressions for the heat and work

    W = -RT ln(V2/V1) = RT ln(P2/P1)

    Q = RT ln(V2/V1) = -RT ln(P2/P1)

    Isobaric Processes:

    For a constant pressure (isobaric) process, we can start from the equations written above in terms of pressure and temperature

    RTdW RdT dP RdTP

    = + =

    p pRTdQ C dT dP C dTP

    = =

    Integrating these, we get

    ( ) ( )2 1 2 1W R T T P V V= =

    2

    1

    T

    p

    T

    Q C dT= We also have, as always,

    2

    1

    2

    1

    T

    v

    T

    T

    p

    T

    U C dT

    H C dT

    =

    =

  • Isochoric Processes:

    For constant volume (isochoric) processes, we can use the equations in terms of T and V to get

    dW = -(RT/V) dV = 0

    dQ = CvdT + (RT/V) dV = CvdT

    From which

    W = 0

    2

    1

    T

    v

    T

    Q C dT= We also have, as always,

    2

    1

    2

    1

    T

    v

    T

    T

    p

    T

    U C dT

    H C dT

    =

    =

    Adiabatic Processes: Ideal gas with constant heat capacity

    For an adiabatic process, by definition no heat enters or leaves the system, so dQ = 0.

    We can therefore take any of the expressions we had above for dQ and set it equal to zero. Doing so gives

    dQ = CvdT + (RT/V) dV = 0, or CvdT = - (RT/V) dV

    and 0, or p pRT RTdQ C dT dP C dT dPP P

    = = =

    and 0, or p pv vC P C PC V C VdQ dP dV dP dV

    R R R R= + = =

    If the heat capacities are constant, we can integrate each of these to get

  • 2 21 1

    2 1

    1 2

    ln ln

    v

    v

    vR

    C

    dT R dVT C V

    T VRT C V

    T VT V

    =

    =

    =

    and

    2 21 1

    2 2

    1 1

    ln ln

    p

    p

    p

    RC

    dT R dPT C P

    T PRT C P

    T PT P

    =

    =

    =

    and

    2 21 1

    2 1

    1 2

    ln ln

    p

    v

    p

    v

    p

    vC

    C

    CdP dVP C V

    CP VP C V

    P VP V

    =

    =

    =

    Often, we use the relationship Cp = Cv + R and we define the heat capacity ratio as

    p pvv v p

    C CC RC C C R

    += = =

    To get

    1

    1 11

    v

    p v

    p p

    RC

    C CRC C

    =

    = = =

    Which then gives

  • ( )

    ( )

    12 1

    1 2

    1

    2 2

    1 1

    2 1

    1 2

    T VT V

    T PT P

    P VP V

    =

    =

    =

    for an ideal gas with constant heat capacity in an adiabatic process

    which we can also write as

    ( )

    ( )1

    1

    constant

    constant

    constant

    TV

    TP

    PV

    =

    =

    =

    for an ideal gas with constant heat capacity in an adiabatic process

    Polytropic Processes:

    We define a polytropic process as one for which

    PV = constant

    For some value of . For this model, SVA give equations for relating T and V, for relating T and P, and for computing heat and work requirements. Rather than derive them, we will now work through examples 3.2 through 3.6 (or as many as possible before the end of class).

  • For gases, we know that at constant temperature, the product of pressure and specific volume is approximately constant. Therefore, we might try to write an equation of state for gases that looks like:

    2 3 ...PV a bP cP dP= + + + +

    where the coefficients a, b, c, etc. are functions only of temperature (and not of pressure). This is an infinite series, but if the PV product is a weak function of pressure, then it can be truncated after a few terms. We could also write it as

    ( )2 31 ' ' ' ...PV a B P C P D P= + + + + where we've simply re-defined the constants as b = aB', etc.

    It turns out that a in the above equation is the same temperature-dependent function for all gases, though the other constants are, in general, different for each particular gas. Experimentally, as the pressure becomes very small (mathematically, in the limit that P goes to zero) the pressure-volume product (PV) goes to the same value for all gases. This is used to establish the ideal gas temperature scale and, in turn, the value of the Universal Gas Constant. If we denote the gas-independent value of PV as the pressure goes to zero as (PV)*, then we can define the ideal gas temperature scale by assigning the quantity (PV)* to be proportional to temperature

    ( )*PV a RT= = where the proportionality constant is the Universal Gas Constant. R is then defined by setting the triple-point of water to be at T = 273.16 K and using the experimental value of (PV)* at the triple-point of water, which is (PV)* = 22711.8 cm3 bar mol-1. This gives

    3 -1 -1 3 -1 -1 -1 -122711.8 83.1447 cm bar mol K 8.31447 m Pa mol K 8.31447 J mol K273.16

    R = = = =

    So, this gives us the first term in our virial equation of state - the pressure (and specific volume) independent term that is the same for all gases. Since it will always be the same, we could move it to the other side of the equation and write

    2 31 ' ' ' ...PV PVZ B P C P D PRT a

    = = + + + +

    We define Z as the compressibility, which is equal to 1 for an ideal gas. Variations in Z away from 1 then indicate non-ideal behavior. Sometimes the virial equation of state is written as an infinite series expansion in inverse volume rather than in pressure, in which case the virial coefficients are different.

    VIRIAL EQUATIONS OF STATE:

  • 2 31 ...PV B C DZRT V V V

    = + + + +

    where, with some mathematical manipulation, we could relate the two forms of the virial coefficients as

    ( )2

    2

    '

    '

    BBRTC BC

    RT

    =

    =

    and so forth.

    In either form (volume-expansion or pressure-expansion), the virial equations will only be useful for practical applications when they converge within a few terms. This is true both because the computations would become inconvenient and because only the first few coefficients are likely to be available from experiments. So, for practical applications, we will truncate these to a one-, two-, or three-term form. The one-term form, in which we keep only the first term is just the ideal gas law:

    1PVZRT

    =

    Truncating the pressure-expansion version after two terms gives

    1 ' 1PV BPZ B PRT RT

    = + = +

    where, in the second version, we have substituted ' BBRT

    = . Alternatively, we could truncate

    volume expansion as

    1PV BZRT V

    = +

    but the other expression in terms of P is more convenient and more commonly used.

    We could also truncate the volume expansion after 3 terms to get a 3-term virial equation

    21PV B CZRT V V

    = + +

    This equation gives P in terms of V directly, but is cubic in V and therefore if one wants to solve for V, one generally does so iteratively. Remember that the coefficients B and C are temperature dependent, and therefore the solution for T in terms of P and V (which we usually don't want to do anyway) may be simple or complex depending on the temperature dependence.

  • Virial coefficients beyond the 3rd virial coefficient are rarely known. If we look at a plot of the compressibility of a typical substance like this one for CO2, which I've prepared from data downloaded from the NIST WebBook, we can get an idea of the range over which a 1, 2, or 3-term virial expansion may be appropriate. A similar plot for methane is given on p. 87 of SVA.

    Here is the same plot with the different truncations of the virial equation superimposed on it.

    Compressibility of CO2

    0.6

    0.7

    0.8

    0.9

    1

    1.1

    1.2

    1.3

    1.4

    0 10 20 30 40 50 60 70 80 90 100Pressure (MPa)

    Z =

    (PV)

    /(RT) 600 K

    400 K

    Compressibility of CO2

    0.6

    0.7

    0.8

    0.9

    1

    1.1

    1.2

    1.3

    1.4

    0 10 20 30 40 50 60 70 80 90 100Pressure (MPa)

    Z =

    (PV)

    /(RT)

    400 K, actual

    600 K, actual

    600 K, 3-term

    600 K, 2-term

    1-term

    400 K, 2-term and 3-term

  • The values of the virial coefficients are obtained from the slope and curvature of the curves as P goes to zero:

    0

    2

    20

    '

    1'2

    P

    P

    dZBdP

    d ZCdP

    =

    =

    For this particular case, we see that at 400 K, the 3-term expansion doesn't work any better than the 2-term expansion, and both are good to about 10 or 15 MPa (100 to 150 atm). At 600K, the 2-term expansion is only good to about 10 MPa, but the 3-term expansion is good out to more than 50 Mpa. Clearly both the values of the virial coefficients and the range over which a given expansion is useful depend on temperature.

    For precise fitting of measured PVT data, one can use an extended virial equation in which additional terms are added to the 3-term virial expansion to allow more flexibility in data fitting. An example is the Benedict/Webb/Rubin (BWR) equation of state:

    2

    0 0 02 3 6 3 2 2 21 exp

    B RT A C TRT bRT a a cPV V V V V T V V

    = + + + + +

    This has 8 parameters, and sometimes some of these are taken to have additional temperature dependence. We will not make much use of the BWR equation of state in this course, but for substances for which the parameters are available, it can provide high accuracy.

    Example 3.7 from SVA

    Cubic Equations of State:

    So far, we have talked about equations of state that are basically good for gases - ideal or non-ideal, but still gas-like (that is, with compressibility reasonably near 1). There are many situations where we would like to have an equation of state that describes both the liquid and the vapor phase of a substance. The simplest and most widely used of such equations are cubic equations of state. By cubic, we just mean that they are cubic polynomials in the specific volume. The first (historically) and simplest of these is the van der Waals equation of state.

    The van der Waals Equation of State:

    This was the first practical cubic equation of state, and it is:

    2RT aP

    V b V=

  • where a and b are positive. If a and b were both zero, this would become the ideal gas equation of state.

    Some general features of cubic equations of state are illustrated for the van der Waals EOS as plotted for CO2 below.

    For this, and all cubic equations of state, there is always a single pressure corresponding to a given volume (at any temperature), but there may be either one or three volumes corresponding to a given pressure. For temperatures above the critical temperature, there is only a single volume for a given pressure (or at least only a single physically meaningful volume, which must be greater than b). For temperatures below the critical temperature, there are 3. The smallest corresponds to the liquid phase, while the largest corresponds to the vapor phase. The middle one doesn't have a direct physical meaning. We will consider this all in more detail next time.

    van der Waals EOS for CO2

    0

    50

    100

    150

    200

    250

    0 200 400 600 800 1000 1200 1400

    Volume (cm3/mol)

    Pres

    sure

    (bar

    )

    600 K

    400 K

    300 K

    250 K

    200 K

  • r

    c

    rc

    TTTPPP

    That is, if we scale the temperature and pressure by the critical temperature and pressure, all fluids are about the same. This is nearly true for simple fluids (monatomic gases except helium), but less nearly true for other substances. The description of other fluids can be improved by introducing a third parameter. The most popular 3rd parameter is the acentric factor, . The acentric factor of a pure substance is defined in terms of its vapor pressure. If we plot the log of the vapor pressure vs. inverse temperature, we get approximately a straight line. If we plot this as reduced pressure vs. inverse reduced temperature, we get a plot like the one shown below.

    of the word theorem. It might better be termed a 'coarse generalization of experimental observations' or something like that. It is not a statement that is mathematically provable or that is always true. The reduced temperature and reduced pressure are defined as

    The Theorem of Corresponding States The two-parameter theorem of corresponding states, which is based on experimental observations, says that all fluids, when compared at the same reduced pressure and temperature, have about the same compressibility. Note that this is an unusual use

  • Vapor Pressure Curves

    -4

    -3.5

    -3

    -2.5

    -2

    -1.5

    -1

    -0.5

    01 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2

    1/Tr (1/K)

    log 1

    0(Pr,s

    at)

    Argon (blue)

    Methane (magenta)

    Hexane

    Ethanol

    Tr=0.7

    The acentric factor is defined as the difference in the value of this curve from the value of -1.0 that is observed at T

    r = 0.7 for simple fluids (argon, neon, krypton). Thus, to find it one needs

    one vapor pressure measurement (at Tr = 0.7) as well as the critical properties. This is the basis

    of the three-parameter theorem of corresponding states: all fluids that have the same acentric factor, when compared at the same reduced pressure and temperature, have about the same compressibility. This is more generally true than is the 2-parameter theorem of corresponding states. That is, it agrees with reality for more cases, because it contains one more parameter and is therefore more flexible. The acentric factor is used in some equations of state, like the Soave-Redlich-Kwong equation of state and the Peng-Robinson equation of state, where it shows up in the function a(T). It is also used in the generalized correlations for the compressibility of gases that we will discuss shortly.

  • Z = Z0 + Z1

    where Z is the compressibility of the gas (or supercritical fluid), is the acentric factor that we just discussed, and Z0 and Z1 are both functions of the reduced temperature and reduced pressure. The most popular of these correlations is the Lee/Kesler correlation. For it, values of Z0 and Z1 are found in appendix E of the textbook. To use these, you first look up the critical temperature, critical pressure, and acentric factor for the substance of interest, then compute the reduced pressure and temperature, then look up Z0 and Z1 in the tables (interpolating between entries) and then finally put them in the above equation for Z. For nonpolar or slightly polar gases, this correlation should give the compressibility to with 2 or 3 percent.

    Generalized correlations for gases: There are a variety of generalized correlations that can be used to estimate the PVT behavior of real (non-ideal) gases. These are often of the form

    hclTypewritten textUnsolved Problems

    prob. 3.6, 3.8, 3.14, 3.27, 3.46

  • Over a limited range of reduced temperatures and pressures (shown on p. 103 of SVA), this correlation can be represented by a two-term truncated virial equation of state as follows. The compressibility is written as

    1 1 c rc r

    BP PBPZRT RT T

    = + = +

    with the correlation

    0 1cc

    BP B BRT

    = +

    which gives

    0 11 r rr r

    P PZ B BT T

    = + +

    or, in the form of the original Lee/Kesler correlation:

    0 1

    0 0

    1 1

    1 rr

    r

    r

    Z Z ZPZ BT

    PZ BT

    = +

    = +

    =

    The coefficients B0 and B1 are functions only of temperature (remember that the virial coefficients themselves are functions only of temperature). They are well represented by:

    01.6

    14.2

    0.4220.083

    0.1720.139

    r

    r

    BT

    BT

    =

    =

    Generalized correlations for liquids The Lee/Kesler correlation discussed above for gases also includes data for subcooled liquids. Figure 3.14 on p. 101 of SVA shows its predictions for both liquids and gases. For saturated liquids (those in equilibrium with the vapor) some simpler correlations work. The Rackett equation gives good results and only requires the critical parameters as inputs:

    ( )0.28571 rTsat

    c cV V Z

    = Another graphical correlation for liquid densities is shown in figure 3.17 on p. 109 of SVA. So, now we have a variety of equations of state and correlations to compute the PVT behavior of fluids. In the homework and in-class examples, we will practice applying these for such direct calculations (i.e. compute volume given temperature and pressure), and then we will go on to use them throughout the rest of the semester to solve more complicated and more interesting problems.

  • and vV T T

    U U UdU dT dV C dT dVT V V

    = + = +

    where we have used the definition of Cv that we have already discussed in previous lectures. The last term (the dV term) is zero for Any constant-volume process, with any fluid Any fluid where internal energy does not depend on molar volume, including

    Ideal gases Incompressible liquids

    In those cases, we have simply

    2

    1

    vT

    v

    T

    dU C dT

    U C dT

    =

    = Similarly for enthalpy, we write H = H (T,P)

    and pP T T

    H H HdH dT dP C dT dPT P P

    = + = +

    where we have used the definition of Cp that we have already discussed. The last term (the dP term) is zero for Any constant-pressure process, with any fluid

    When we put heat into (or remove heat from) a fluid, this heat might be used (in whole or in part) to do work on the surroundings, to change the temperature of the fluid, to change the phase of the fluid, or to provide the energy required to carry out a chemical reaction. In the next couple of lectures, we will talk about these heat effects. The first ones that we will consider are sensible heat effects, which involve heating that changes the temperature of the system. Sensible Heat Effects If we consider a closed system where there are no phase transitions, no chemical reactions, and no changes in the composition of the system, then adding heat to or removing heat from the system will change its temperature and/or cause it to do work on the surroundings. Our goal here is to relate the temperature change and work done to the amount of heat added. For a homogeneous substance of constant composition, the phase rule tells us that two intensive properties must be fixed to specify the state of the fluid completely. Thus the internal energy or enthalpy can be written as a function of two intensive variables, like temperature, pressure, and molar volume. Thus, using T and V as the independent variables for U, we write U = U (T,V)

    hclTypewritten textSENSIBLE HEAT AND LATENT HEAT

  • Any fluid where the enthalpy does not depend on pressure, particularly ideal gases In those cases, we have simply

    2

    1

    p

    T

    p

    T

    dH C dT

    H C dT

    =

    = Furthermore, we know that for constant-volume processes, Q = U, and for constant-pressure processes, Q = H. This lets us calculate things like the amount of heat required to change the temperature of a substance by a certain amount. Heat capacities are often expressed in the form of polynomials in temperature, or as other simple functions. The one used in SVA most commonly is

    2 2pC

    A BT CT DTR

    = + + +

    Note that for ideal gases, the constant volume and constant pressure heat capacities are related by Cp = Cv + R So for an ideal gas, if Cp is given by the form shown above, then Cv is given by

    2 21vC A BT CT DTR

    = + + +

    For incompressible liquids, Cv and Cp are the same (a constant pressure process on an incompressible substance is also a constant-volume process, since the volume is always constant for an incompressible substance). The above form of the heat capacity is easily integrated - for example

    ( )

    ( ) ( ) ( )

    2 2

    1 1

    2 2

    2 2 3 32 1 2 1 2 1

    2 1

    1 12 3

    T T

    p

    T T

    H C dT R A BT CT DT dT

    B CH R A T T T T T T DT T

    = = + + +

    = + +

    ,all

    p i p ii

    C y C=

    SVA gives a longer description of evaluation of this, but I assume you can all do integrals, so let's not spend time on it. They also give a representative computer program for doing this in appendix D. We might similarly have a handy function in an Excel spreadsheet For mixtures of gases at low pressure, one can simply take the heat capacity of the mixture as the average of the heat capacity of the components, weighted by their mole fractions in the mixture. That is, if y

    i is the mole fraction of species i in the mixture, then the molar heat capacity of the

    mixture is simply

  • Latent Heats of Pure Substances When a pure substance is converted from one phase to another (solid to liquid, liquid to vapor, etc.) heat flows to or from the substance, but there is no change of temperature. The energy per mole required to vaporize a liquid is called the latent heat of vaporization, and the energy per mole required to melt a solid is called the latent heat of fusion. We know from the phase rule that for a 1-component, 2-phase system, specifying one intensive variable specifies the state of the system. Thus, the latent heat of a phase change is a function only of one intensive variable, and can therefore be written as a function of temperature alone. This function can be written as

    satdPH T V

    dT =

    where H is the latent heat for the phase change (the difference between the molar enthalpies of the two phases), V is the volume change for the phase change (big for vaporization, small for melting), and Psat is the pressure of the two-phase system (which is a function only of temperature, since specifying the temperature specifies all other intensive properties of the 2-phase system). For a vaporization phase change, Psat is the vapor pressure of the liquid. This equation, which we will derive in a few weeks, is called the Clapeyron equation. So, to get the heat of vaporization, we could either make a direct measurement (using a calorimeter), or we could compute it from the vapor pressure curve and the volumetric properties of the liquid and gas. Some correlations for estimating heats of vaporization are given on p. 131 of SVA. I won't bother to repeat them here, but clearly you should be able to apply them.

  • The process is carried out at constant pressure, and no non-flow work is done, so Q = H That is, the heat added or removed is equal to the change in enthalpy of the stream flowing through the calorimeter. This, in turn, is equal to the difference in enthalpy between the products and the reactants, at constant temperature. The heat of reaction for the reaction is defined as this change in enthalpy at constant temperature. Now, consider the generic balanced chemical reaction

    aA + bB lL + mM The standard heat of reaction for this reaction at temperature T is defined as the enthalpy change when a moles of A plus b moles of B in their standard states at temperature T react to form l moles of L and m moles of M in their standard states at the same temperature T. A standard state is a particular state of a species at temperature T and at specified conditions of pressure, composition, and physical condition (e.g. gas, liquid, or solid). In SVA, the standard state conditions for computing heats of reaction are: Gases: the pure substance in the ideal-gas state at 1 bar. Liquids and solids: the pure substance at 1 bar. Older data tabulations often use 1 atm rather than 1 bar as the standard pressure, but this makes a negligible difference for most applications. It is particularly important to note and remember

    Reactants in at To Products out at To

    Q (in or out)

    Isothermal Flow

    Calorimeter

    Heats of Reaction So far, the heat effects we have discussed (sensible heat and latent heat of pure substances) have been for physical processes (processes where no chemical reactions occur). In fact, pretty much everything that we have done so far this semester has been for physical processes. However, as you might guess, chemically reacting systems are an important part of chemical engineering. Heat effects due to chemical reaction can be quite large compared to those associated with physical processes. During a chemical process, energy is stored in and released from the chemical bonds that form between atoms in molecules. Since these bonds are, usually, much stronger than the physical interactions between molecules, they can store and release much greater amounts of energy. Consider an isothermal flow calorimeter for measuring the heat released or consumed by reaction. In such a device, a stoichiometric mixture of reactants flows in and reacts, and we measure the heat (addition or removal) required to return the reaction products to the inlet temperature of the reactants.

  • that the heat of reaction is defined for a particular set of stoichiometric coefficients of the reaction. For example, we could write the reaction for the partial oxidation of methane to CO and H2 (synthesis gas) as CH4 + O2 CO + 2 H2 or as 2 CH4 + O2 2 CO + 4 H2 Either of these is a valid way to write the reaction, but the heat of reaction for the second version is twice that of the first version, so it is important that one know which version is being used. A heat of reaction can only be interpreted with respect to a particular writing of the reaction. Standard Heats of Formation: It would be impossible (or at least impractical) to tabulate the heats of reaction for all possible reactions of interest. Therefore, they are calculated from a standard set of heats of reaction. This set is the heats of formation of the chemical species. The heat of formation of a substance is defined as the heat of reaction when the substance is formed, in its standard state at temperature T, from the elements in their standard states at temperature T. For example, the heat of formation of ethanol (CH3CH2OH) is equal to the heat of reaction for the reaction 2 C(s) + 3 H2(g) + O2(g) CH3CH2OH(l) This is the formation reaction for ethanol. In this hypothetical reaction, two moles of graphite (the standard state of carbon) plus 3 moles of hydrogen gas plus one-half of a mole of oxygen gas produce one mole of liquid ethanol. Heats of formation are defined as the heat of reaction per mole of the species formed, and therefore these reactions are written so that the stoichiometric coefficient of the species formed is one. The value of heats of formation is that we can write any reaction as the sum of formation reactions. Consider again the partial oxidation of methane to give synthesis gas CH4 + O2 CO + 2 H2 This can be written as the sum of the following formation reactions: C + O2 CO CH4 C + 2 H2 CH4 + O2 CO + 2 H2 Note that we don't have to include formation reactions for H2 or O2, since they are elements. The heat of reaction of the first reaction (C + O2 CO) is the heat of formation of CO (Hf(CO)). The heat of reaction of the second reaction is minus the heat of formation of CH4 (-Hf(CH4)) (reversing the direction of the reaction reverses the sign of the heat of reaction). So, the heat of reaction for the partial oxidation is Hrxn = Hf(CO) - Hf(CH4). In general, any heat of reaction can be computed as the heat of formation of the products minus the heat of formation of the reactants. That is, for the generic reaction

    aA + bB lL + mM the heat of reaction is

    Hrxn = lHf(L) + mHf(M) - aHf(A) - bHf(B)

  • Thus, if we have a table of the enthalpies of formation of species (at some standard temperature), then we can compute the heat of reaction for any reaction involving those species (at the same standard temperature). The heats of formation of elements are zero by definition. They form the reference point for defining the heats of formation of everything else. Usually, heats of formation are tabulated at a standard temperature of 298 or 298.15 K. To get heats of formation or heats of reaction at other temperatures, we must integrate the heat capacities of the reactants and products to see how the enthalpies of the reactions and products change with temperature. Heats of Combustion: Most formation reactions cannot actually be carried out in practice. For example, we probably could not devise a calorimeter where we form ethanol (in 100% yield) from a stoichiometric mixture of graphite, hydrogen, and oxygen. So, heats of formation usually have to be determined from calorimetric measurements that can easily be carried out. The most common class of reactions that goes nicely to completion is combustion reactions. Imagine that we have a flow calorimeter in which we can burn things. Then we could first burn some hydrogen H2(g) + O2(g) H2O(l) From which we could get the heat of formation of water (this is a formation reaction that we can measure directly). Next, we might be able to burn some graphite directly, as C(s) + O2(g) CO2(g) From which we would get the heat of formation of CO2 (this is another formation reaction that we can measure directly). Knowing these, we could boldly start burning all sorts of compounds made up of only C, H, and O, and could determine their heats of formation from the heats of reaction for the combustion reaction CxHyOz + (x + y - z) O2(g) xCO2(g) + y H2O(l) and the known heats of formation for carbon dioxide and water. This particular heat of reaction (to convert 1 mole of the substance burned plus a stoichiometric amount of oxygen to carbon dioxide and water) is known as the heat of combustion of the material. For example, we could burn ethanol as CH3CH2OH(l) + 3 O2(g) 2 CO2(g) + 2 H2O(l) The standard heat of reaction for this reaction at 298 K is ( ) ( ) ( ),298 ,298 2 ,298 2 ,298 3 22 ( ) 2 ( ) ( )o o o orxn f f fH H CO g H H O l H CH CH OH l = + This is the heat of combustion for ethanol. So, we might also write it as

    ( ) ( ) ( ) ( ),298 3 2 ,298 2 ,298 2 ,298 3 2( ) 2 ( ) 2 ( ) ( )o o o ocomb f f fH CH CH OH l H CO g H H O l H CH CH OH l = + Note that in these reactions we have explicitly indicated the phase (s for solid, l for liquid, or g for gas) that is being taken as the standard state for each species. From the above From the above relationship between the heat of formation and the heat of combustion, we have

    ( ) ( ) ( ) ( ),298 3 2 ,298 2 ,298 2 ,298 3 2( ) 2 ( ) 2 ( ) ( )o o o of f f combH CH CH OH l H CO g H H O l H CH CH OH l = + So, let's imagine for a moment that we don't know the heats of formation of anything, but we have a flow calorimeter that we operate at a constant pressure of 1 bar and an inlet and outlet temperature of 25 C. We first burn H2 in it, and find a heat of combustion of -68.3174 kcal/mol. Next, we burn some graphite and find a heat of combustion of -94.0518 kcal/mol. Finally, we

  • burn some ethanol and find a heat of combustion of -258.3884 kcal/mol. The formation reactions for H2O and CO2 are the same as the combustion reactions of H2 and C, respectively, so the heats of formation for H2O and CO2 are equal to the heats of reaction that we measure for the H2 and C combustion reactions (both are at the same standard state conditions of 25 C and 1 bar). So, we get

    ( )( )

    ,298 2

    ,298 2

    ( ) 95.0518 kcal/mol

    ( ) 68.3174 kcal/mol

    of

    of

    H CO g

    H H O l

    =

    =

    Then, we can compute the heat of formation of ethanol from these, and from the heat of combustion that we measured for it:

    ( ) ( ) ( ) ( )( )

    ,298 3 2

    ,298 3 2

    ( ) 2 94.0518 2 68.3174 267.6988 kcal/mol

    ( ) 66.35 kcal/mol = -277.61 kJ/mol

    of

    of

    H CH CH OH l

    H CH CH OH l

    = +

    =

    Reassuringly, this agrees with the value in the table in the back of SVA to within 0.1 kJ/mol. I took the data for the example from Perry's Handbook. The Temperature Dependence of Heats of Reaction: So, now we know how to use heats of formation to compute heats of reaction at a given temperature, but what if we want to know a heat of reaction at some temperature other than the standard temperature of 25 C at which standard heats of formation are generally tabulated? One way that we could compute this would be to compute the enthalpy change for a hypothetical process in which we heat or cool the reactants to the standard state temperature, then carry out the reaction isothermally, then cool or heat the products back to the temperature of interest. The total enthalpy change for this 3-step process would be equal to the entropy change for the 1-step process of carrying out the reaction at the initial temperature. For the first step of heating or cooling the reactants to 298.15 K, we have

    298.15 K

    total1 ,reactantso

    pT

    H C dT = where total,reactantspC is the total heat capacity of the reactants for the reaction as written. For example, if we had the generic reaction

    aA + bB lL + mM it would be total,reactants ( ) ( )p p pC aC A bC B= + For the second step, carrying out the reaction at 298.15 K, we have 2 ,298.15 Ko orxnH H = Finally, for the 3rd step, bringing the products back to the initial temperature, we have

    total3 ,products298.15 K

    To

    pH C dT = So, for the overall process, we have

    298.15 K

    total total, ,reactants ,298.15 K ,products

    298.15 K

    To o ooverall rxn T p rxn p

    TH H C dT H C dT = = + +

    Remembering that switching the limits of an integral (changing the direction of integration) just changes its sign, we can combine the two integrals to get

  • ( )total total, ,298.15 K ,products ,reactants298.15 K

    To orxn T rxn p pH H C C dT = +

    That is, the heat of reaction at some other temperature is equal to the heat of reaction at the standard temperature of 298.15 K plus the integral from the standard temperature to the temperature of interest of the difference in heat capacity between the reactants and the products. Just like the enthalpy of reaction is the difference between the enthalpy of the products and reactants, we can define the change in heat capacity of reaction as the difference between the total heat capacity of the products and the reactants. So, for the generic reaction

    aA + bB lL + mM For which the heat of reaction is

    Hrxn = lHf(L) + mHf(M) - aHf(A) - bHf(B) The difference in heat capacity on reaction would be

    Cp = lCp(L) + mCp(M) - aCp(A) - bCp(B) where all of the heat capacities can be temperature dependent, so their difference is also temperature dependent. The heat of reaction at some arbitrary temperature T is then

    , ,298.15 K298.15 K

    To orxn T rxn pH H C dT = +

    We could further generalize this by writing the general chemical reaction with an arbitrary number of reactants and products as done by SVA on p. 137 1 1 2 2 3 3 4 4... ...A A A A + + + + where i is the stoichiometric coefficient of species Ai. The Ai (A1, A2, etc.) are just the names of the species participating in the reaction. The convention we will use here (and that most widely used) is that the stoichiometric coefficients are positive for products and negative for reactants. Then, the above equation can be formally written as: 1 1 2 2 3 3 4 4 ... 0A A A A + + + + = or simply

    0i ii

    v A = With this notation, the heat of reaction can be written as

    ( )o orxn i f ii

    H v H A = The change in heat capacity can be written as

    ,p i p ii

    C v C = and so on. This is very handy for doing things in an automated way on the computer. If the individual heat capacities of the species are written in the usual polynomial form that we used last time

    , 2 2p i i i i iC

    A B T C T D TR

    = + + +

    Then we can write

  • ( )2 2,

    2 2

    p i p i i i i i ii i

    p

    C v C R v A B T C T D T

    CA BT CT DT

    R

    = = + + +

    = + + +

    where we define

    , , , i i i i i i i ii i i i

    A v A B v B C v C D v D = = = = Using this notation with the heat capacity integral to find the heat of reaction at some arbitrary temperature T, we have

    ( )( ) ( )( ) ( )( )

    , ,298.15 K298.15 K

    2 2, ,298.15 K

    298.15 K

    2 32 3, ,298.15 K

    1 1298.15 K 298.15 K 298.15 K2 3 298.15 K

    To orxn T rxn p

    To orxn T rxn

    o orxn T rxn

    H H C dT

    H H R A BT CT DT dT

    B CH H R A T T T DT

    = +

    = + + + +

    = + + +

    As was the case for the usual heat capacity integral, SVA have given a defined function and an example computer program for this in the appendices of the text. Similarly, we might use a spreadsheet . This example spreadsheet is set up to evaluate the heat of reaction for methanol synthesis from synthesis gas, as shown in example 4.6 in SVA. The final section in chapter 4 of SVA is on 'Heat Effects of Industrial Reactions', which we will study by working a couple of the examples from that section, which also illustrate the other things we've been looking at today.

    hclTypewritten textUnsolved Problems

    Prob No. 4.8 and 4.19

  • The first law of thermodynamics says that energy is conserved, and does not differentiate between energy added to or removed from a system as heat and energy added to or removed from a system as work. In the first law, heat and work are equivalent. However, our everyday experience, as well as centuries of scientific studies and engineering experience shows that heat and work are not equivalent in all ways. Work is spontaneously converted to heat all around us, and we can find many examples where work is fully converted to heat. The classical example of this is the historic experiments of Joule where he stirred fluids and measured the amount of work done and the temperature rise of the fluids. However, there are no examples of energy in the form of heat being spontaneously and completely converted to work, despite centuries of effort directed at constructing a device that would do so. This suggests that there is some sense in which work is a higher or more valuable form of energy than heat, and in fact if that were not the case, there would not be much need to differentiate between the two. This idea is formalized and quantified in the second law of thermodynamics.

    There are many ways to state the second law of thermodynamics. Given one of them, it is possible (sometimes by simple means, sometimes by more elaborate means) to derive the others. Two possible statements of it (from SVA, p. 156) are:

    No device can operate such that its only effect is to convert heat absorbed by a system completely into work done by the system.

    Or

    No process is possible that consists solely of the transfer of heat from one temperature level to a higher one. We will provide more mathematical statements of the second law shortly. This will all be done quickly, on the assumption that it is review of material recently learned in EAS 204. Heat Engines, Carnot Engines, and so forth: A heat engine is a device that converts heat to work. Much of classical thermodynamics was developed in the context of trying to understand and improve steam engines, which powered locomotives, steamboats, etc. in times past. In these and in other heat engines, heat is absorbed by the system at some high temperature and partially converted to work and partially transferred out of the system at some lower temperature. For example, in a steam engine, water flows into a boiler where it is vaporized and superheated (absorbing heat at a high temperature), then the steam expands through a mechanical device like a turbine, converting part of this heat to work. The steam then condenses (transferring heat out of the system at a low temperature) and is returned to the boiler, so that the overall process is cyclic. The working fluid of the heat engine (i.e. the steam) absorbs some heat |QH| at a high temperature, does some net amount of work |W| on the surroundings, then rejects some heat |QC| at low temperature. Since the overall process is

    THE SECOND LAW OF THERMODYNAMICS

  • cyclic, the overall change in the state of the system (and therefore in the internal energy of the working fluid) is zero. So, the first law tells us that |W| = |QH| - |QC| That is, the net work done is equal to the difference between the heat flow in at high temperature and the heat flow out at low temperature. Note that for the example of the steam engine, the net work is some relatively large amount of work done by the expanding steam minus some relatively small amount of work that must be done to pump the water up to the high pressure of the boiler. The thermal efficiency of a heat engine is simply defined as the fraction of the heat input that is converted to work. That is

    net work output 1heat absorbed

    H C C

    H H H

    W Q Q QQ Q Q

    = = =

    If we could just eliminate the rejection of heat at from the system (make |QC| zero), we could achieve unity (100%) efficiency. However, the second law of thermodynamics forbids this. In fact, the most efficient possible engine is one that operates in an entirely reversible fashion (with no irreversibilities). Such an idealized engine is called a Carnot engine, after N.L.S. Carnot, who first described it. In such a (hypothetical) device, the following four steps make up the engine cycle: 1. Reversible, adiabatic compression from temperature TC to temperature TH. 2. Reversible, isothermal heat transfer during which |QH| is absorbed from the surroundings. 3. Reversible, adiabatic expansion from temperature TH to TC. 4. Reversible, isothermal heat transfer during which |QC| is rejected to the surroundings,

    bringing the working fluid back to its original state. Any engine that operates in a fully reversible way is considered a Carnot engine. In SVA, they demonstrate that (a) no engine can have a higher thermal efficiency than a Carnot engine (for the same high and low temperature levels) and (b) the thermal efficiency of a Carnot engine depends only on the high and low temperature levels. Thermodynamic Temperature Scales: The fact that the efficiency of a Carnot engine depends only on temperatures and not on any properties of the working fluid allows us to establish a thermodynamic temperature scale that is independent of any material properties. Conveniently, this will work out to be the same as the ideal gas temperature scale (and the Kelvin scale) that we have already been using. Assuming that this was well covered in EAS204, we will skip the derivations. In SVA, they demonstrate that if temperature is defined by the ideal gas temperature scale, then the thermal efficiency of a Carnot engine is

    1 CH H

    W TQ T

    = =

    and the heat absorbed and heat rejected are related by

    H HC C

    Q TQ T

    =

    To get 100% efficiency, we would have to either have a cold reservoir at 0K, or a hot reservoir at infinite temperature. Obviously, neither of these is achievable.

  • Entropy: From the above discussion of Carnot engines, we get (among other things) a relationship between the heat absorbed at TH and the heat rejected at TC in a Carnot cycle:

    H HC C

    Q TQ T

    =

    or

    H CH C

    Q QT T

    =

    If we take the working fluid as our system and use our convention that heat flow into the system is positive, then QH is positive and QC is negative, and we have

    CHH C

    QQT T

    =

    or

    0CHH C

    QQT T

    + =

    Thus, for a complete cycle, the two quantities Q/T associated with heat transfer into and out of the system sum to zero. A key characteristic of a thermodynamic property (like temperature, pressure, internal energy, etc.) is that for any cyclic process, its final value is the same as its initial value. This suggests that there may be another thermodynamic property of the system whose changes are given by Q/T. In SVA they approach this by pointing out that an arbitrary reversible, cyclic process can be constructed from many Carnot cycles in which an infinitesimal amount of heat dQH is absorbed at temperature TH and another infinitesimal (but different) amount of heat dQC is rejected at temperature TC. From such an analysis, one can conclude that

    0revdQT

    =v That is, the integral over any cyclic, reversible process of dQ/T is zero. This is then used to define a new property entropy, denoted by the symbol St, as

    t revdQdST

    =

    or trevdQ TdS= The superscript t is used here to indicate that this is the total entropy of the system (an extensive quantity) not the molar or specific entropy (an intensive quantity). We can show that entropy is a state function, since we could evaluate the change in entropy from state A to state B by doing the integral

    1A to B via path 1

    t revdQST

    = along one reversible path, and then do the integral from state B back to A along another reversible path

    2B to A via path 2 A to B via path 2

    t rev revdQ dQST T

    = = Since the total integral over the cyclic path from A to B and back to A must be zero, we have

  • 1 2 0t tS S + = Since we are free to pick an arbritrary reversible path for either part 1 or part 2, the entropy change must be independent of the path selected, and

    1 2A to B by any reversible path

    t t t trevB A

    dQS S S ST

    = = = If the system goes from state A to state B by an irreversible process, then the entropy change will still be the same, but it will not be given by the integral, of dQ/T along the irreversible path. So, even for irreversible processes, we can always compute the entropy change along some hypothetical reversible path that connects the same initial and final states as the actual irreversible path. If a process is both reversible and adiabatic, then dQrev will be zero throughout the process, and dSt will be zero throughout, and finally, St will be zero. So, a reversible, adiabatic process is also called an isentropic process. The discussion of entropy so far can be summarized as: There exists a property that we will call entropy and denote by S. It is related to observable properties of the system. For a reversible process, changes in this property are given by

    t revdQdST

    =

    The change in entropy for any process that takes a system from state A to B (whether the actual process is reversible or not) is given by

    A to B by any reversible path

    t revdQST

    = Mechanically reversible processes are processes where all driving forces for change in the state of the system are infinitesimal except for temperature differences for heat transfer between the system and the surroundings. For mechanically reversible processes, the entropy change can still

    be calculated as t dQST

    = where T is the temperature of the system (which must be uniform). In this case, the irreversibilities occur outside the system, so from the point of view of the system, it is immaterial what temperature the heat finally flows to. The process is internally reversible - there are no irreversibilities (no finite driving forces) within the system. Entropy Changes for an Ideal Gas To get started with computing entropy changes, we will consider entropy changes for processes involving ideal gasses. For one mole of fluid undergoing a mechanically reversible process in a closed system, the first law tells us that: dU = dQrev - PdV Since (by differentiation of H = U + PV) enthalpy is related to internal energy by dH = dU + PdV + VdP

  • we see that for a mechanically reversible process dH = dQrev - PdV + PdV + VdP = dQrev + VdP or dQrev = dH - VdP The above is true in general (not just for an ideal gas), but if we consider the particular case of one mole of an ideal gas, we have dH = CpdT and V = RT/P, so dQrev = CpdT - RT/PdP so

    rev pdQ dT dPC R

    T T P=

    and therefore

    pdT dPdS C RT P

    =

    Integrating this from some initial state T0, P0 to some final state T and P gives

    0

    0ln

    T

    p

    T

    dT PS C RT P

    =

    Although we used a mechanically reversible process to derive this, the final two equations relate only properties of the system (state functions) and are therefore independent of path and apply to all processes, reversible or irreversible. For an ideal gas with constant heat capacity, the above equation is even simpler

    0 0

    ln lnpT PS C RT P

    =

    For a gas in which the heat capacity is defined by the sort of polynomial that we have been using:

    2 2pC

    A BT CT DTR

    = + + +

    then

    ( ) ( ) ( )0

    30

    2 2 2 20 0 0

    0 0

    ln

    ln ln2 2

    T

    T

    S A D PB CT dTR T PT

    S T C D PA B T T T T T TR T P

    = + + +

    = + +

    ( ) ( ) ( )0

    2 2 2 20 0 0

    0ln

    2 2

    Tp

    T

    C T C DICPS dT A B T T T T T TRT T

    = + +

    As they did for enthalpy and internal energy changes, SVA also express this in terms of a mean heat capacity and provide a defined function and example computer program for it in the appendices. We can also do this calculation in a spreadsheet that evaluates the first part of this:

  • Stotal 0 for all processes That is, the total entropy of the universe (the system plus the surroundings) cannot decrease. For reversible processes, it does not change. For irreversible processes, it increases. This is, in essence, a statement that things always happen in a certain direction. Heat always flows from hot to cold because that increases the total entropy. Or, from a different perspective, we define entropy such that its total value increases when heat flows from hot to cold. The fact that spontaneous heat transfer always occurs from hot to cold is then consistent with the version of the second law as written above. If heat leaves a hot object (at TH) and flows into a cooler object (at TC) then the entropy of the hot object decreases by Q/TH while that of the cooler object increases by Q/TC. So, the total change of entropy of the universe for this process is Q(1/TC 1/TH), which is positive. If the temperatures were only infinitesimally different, the process would be reversible, and this difference would approach zero. Notice that while the first law of thermodynamics is an equality stating that the total amount of energy in the system plus the surroundings remains constant the second law of thermodynamics is an inequality stating that the total amount of entropy in the system plus the surroundings can only remain constant or increase. Entropy balances for open systems: We can write an entropy balance for an open system in the same way that we wrote an energy balance and a mass balance except that entropy is not conserved so we must include generation of entropy as well as its flow in and out of the system. Mathematically, we can write the entropy balance as:

    ( ) ( ) 0tsurrcv

    gfs

    d mS dSSm Sdt dt

    + + = The first term is the net rate of entropy transport out of the system via flow. It is the difference between the entropy carried out by streams flowing out of the control volume and that carried in by streams flowing into the control volume. The second term is the rate of accumulation of entropy within the control volume. The third term is the rate of change of the entropy of the surroundings due to the process of interest. gS is the total rate of entropy generation. It is zero for reversible processes and positive for irreversible processes. If the surface of the control volume is made up of a number of areas through each of which an amount of heat Qj flows from the surroundings at temperature Tj, then the rate of entropy change of the surroundings due to this heat flow will be:

    The Second Law of Thermodynamics A mathematical statement of the 2

    nd

    law of thermodynamics: Given the definition of entropy that we developed/adopted last time, either of the two statements of the second law of thermodynamics given there can be shown to be equivalent to:

  • t

    jsurr

    j j

    QdSdt T

    = Substituting this, the entropy balance becomes

    ( ) ( ) 0jcv gfsj j

    d mS QSm S

    dt T + =

    If the system is at steady state, this simplifies to

    ( ) 0j gfsj j

    QSm S

    T =

    If there is a single inlet and a single outlet from the system, with equal mass flowrates, we can write

    0jout in gj j

    QS S S

    T =

    In this final equation, the heat flow and entropy generation terms are per unit mass of material flowing through the system, rather than the total heat flows and total entropy generation used in the previous equations. Examples 5.5 and 5.6 from SVA

    hclTypewritten textUnsolved Problems

    Prob 5.8, 5.10, 5.12, 5.19, 5.29, 5.35 and others, u can solve for practice

  • The reversible work done on the closed system is dWrev = -Pd(nV) which is our usual expression developed in the first chapter of SVA. The reversible heat flow is (as developed in the previous lecture and chapter 5 of SVA) dQrev = Td(nS) Combining these last 3 equations gives d(nU) = Td(nS) - Pd(nV) or dUt = TdSt - PdVt These last two equations give a general relationship between the primary thermodynamic variables for the fluid - U, S, T, P and V. Although we derived it for a reversible process, the final expressions relate only state variables that are path-independent. Therefore, this relationship holds for any process occurring in a closed system between equilibrium states. The system can be made up of multiple phases, and it can undergo reactions, as long as no material enters or leaves it and the initial and final states that we wish to relate are equilibrium states. For convenience, we define three more thermodynamic properties related to the internal energy. These are the enthalpy (which we have been using) H U PV + the Helmholtz energy (or Helmholtz free energy) A U TS and the Gibbs energy (or Gibbs free energy) G H TS

    In lectures 3 through 6 of this course, we looked fairly intensively at how to compute the volumetric properties (i.e. volume given T and P) for fluids from equations of state and from correlations based on the theorem of corresponding states. After that, we looked at heat effects, and computed enthalpy and internal energy changes when the temperature of a substance is changed using the heat capacity. Then we defined entropy, and computed entropy changes when the temperature of a substance was changed, again using heat capacity data. Now, we want to bring this all together to see how, in general, not just for ideal gases, we can compute changes in the thermodynamic properties of a fluid (internal energy, enthalpy, and entropy) for a change in state of the fluid, using PVT and heat capacity data. We will also define and use two new thermodynamic quantities, the Helmholtz free energy and the Gibbs free energy. Property Relations for Homogeneous Phases First, we will combine the first and second laws to derive a fundamental property relationship between U, S, T, P and V - all of the primary thermodynamic variables. We can write the first law for a closed system containing n moles of fluid as: d(nU) = dQ + dW For the special case of a reversible process, this is d(nU) = dQ

    rev + dW

    rev

    hclTypewritten textTHERMODYNAMIC PROPERTIES OF FLUIDS,

  • For each of these, we can transform the thermodynamic property relation for internal energy using the definitions above to get a new thermodynamic property relation where internal energy is replaced by enthalpy, Helmholtz energy, or Gibbs energy. For enthalpy, we have nH = nU + nPV so d(nH) = d(nU) + Pd(nV) + nVdP or d(nU) = d(nH) - Pd(nV) - nVdP equating this to the fundamental thermodynamic property relationship for d(nU) d(nU) = d(nH) - Pd(nV) - nVdP = Td(nS) - Pd(nV) so d(nH) = Td(nS) + nVdP or dHt = TdSt + VtdP Likewise, differentiating the definition of A gives d(nA) = d(nU) - Td(nS) - nSdT so d(nU) = d(nA) + Td(nS) + nSdT = Td(nS) - Pd(nV) so d(nA) = -nSdT - Pd(nV) or dAt = -StdT - PdVt Finally, for the Gibbs energy, d(nG) = d(nH) - Td(nS) - nSdT so d(nH) = d(nG) +Td(nS) + nSdT = Td(nS) +nVdP so d(nG) = -nSdT + nVdP or dGt = -StdT + VtdP The above are all written for the total mass of a closed system. What we are often interested in is the corresponding relationsh