Thermodynamics and Chemistry

1

Physical Quantities, Units, and Symbols

1.1 Physical Quantities, Units, and Symbols

🧭 Overview

🧠 One-sentence thesis

The International System of Units (SI) provides a standardized framework for expressing physical quantities in thermodynamics, built on seven base units that allow consistent quantitative description of all physical phenomena.

📌 Key points (3–5)

  • What SI provides: A globally agreed system of units for science and technology, based on seven independent base units.
  • Base units cover: time (second), length (meter), mass (kilogram), temperature (kelvin), amount of substance (mole), electric current (ampere), and luminous intensity (candela).
  • Amount of substance: A counting quantity for specified elementary entities (atoms, molecules, ions, etc.), measured in moles.
  • Symbol standardization: IUPAC Green Book recommendations ensure consistent notation across scientific work.
  • Common confusion: "Amount of substance" is a formal term for what is measured in moles—it counts entities, not mass or volume.

📏 The SI system foundation

📏 What the SI is

International System of Units (SI): the internationally agreed system of units for physical quantities in science and technology.

  • SI stands for the French Système International d'Unités.
  • The excerpt emphasizes that there is "international agreement" that SI units should be used in science applications.
  • This standardization allows scientists worldwide to communicate measurements unambiguously.

🔢 The seven base units

The SI rests on seven independent physical quantities, each with its own base unit:

Physical QuantitySI UnitSymbol
timeseconds
lengthmeterm
masskilogramkg
thermodynamic temperaturekelvinK
amount of substancemolemol
electric currentampereA
luminous intensitycandelacd
  • These seven quantities are "independent" and "sufficient to describe all other physical quantities."
  • The candela is noted as "usually not needed in thermodynamics."
  • Definitions of these base units are referenced to an appendix in the source material.

🧪 Amount of substance

🧪 What "amount of substance" means

Amount of substance: a counting quantity for specified elementary entities.

  • This is the formal name for what the mole measures.
  • It is not mass, volume, or concentration—it is a count of entities.
  • Elementary entities can be atoms, molecules, ions, electrons, or "any other particle or specified group of particles."

🔬 The mole as a base unit

  • The mole is the SI base unit for amount of substance.
  • The excerpt emphasizes that you must specify what you are counting (e.g., molecules of water, ions of sodium).
  • Example: One mole of water molecules is a different quantity than one mole of hydrogen atoms, even though both use the same unit.

Don't confuse: Amount of substance (moles) with mass (kilograms) or volume (cubic meters). The mole counts entities; mass and volume measure different properties.

📖 Symbol standardization

📖 IUPAC Green Book

  • The excerpt references the third edition of the IUPAC Green Book as the source for recommended symbols.
  • IUPAC stands for International Union of Pure and Applied Chemistry.
  • The Green Book (named for its cover color) is a "manual of recommended symbols and terminology based on the SI."

📋 Why standardization matters

  • Consistent symbols prevent confusion when reading scientific literature.
  • The source material includes appendices listing physical quantity symbols and abbreviations for convenient reference.
  • The excerpt notes "with a few exceptions" the book follows IUPAC recommendations, showing that even standardization allows some flexibility for clarity.

🔧 Practical context

🔧 Thermodynamics as quantitative

  • The chapter opens by stating "Thermodynamics is a quantitative subject."
  • Physical quantities allow derivation of relations between values.
  • Most quantities require units; only dimensionless quantities (like mole fraction) are "pure numbers."

🔧 What the chapter covers

The excerpt is a table of contents entry, indicating the chapter will discuss:

  • The International System of Units (covered above)
  • Amount of substance and amount (covered above)
  • The SI revision of 2019 (not detailed in this excerpt)
  • Derived units and prefixes (not detailed in this excerpt)
  • Quantity calculus (mentioned but not explained)
  • Dimensional analysis (mentioned but not explained)

Note: The excerpt consists primarily of a table of contents with brief introductory remarks. The substantive content focuses on introducing SI base units and the concept of amount of substance. Topics like quantity calculus, dimensional analysis, derived units, and the 2019 SI revision are listed but not explained in the provided text.

2

Quantity Calculus

1.2 Quantity Calculus

🧭 Overview

🧠 One-sentence thesis

Quantity calculus treats physical quantities as products of numerical values and units, enabling algebraic manipulation that ensures dimensional consistency and simplifies unit conversions, especially within the coherent SI system.

📌 Key points (3–5)

  • Core principle: A physical quantity equals a numerical value multiplied by units; the symbol for a quantity does not imply specific units.
  • Key manipulation: Dividing a physical quantity by its units yields a pure number representing "how many of those units."
  • SI coherence advantage: Using SI base and derived units without prefixes makes all conversion factors equal to unity, simplifying calculations.
  • Common confusion: The symbol (e.g., ρ for density or n for amount) is not "the number of units"—only the ratio (quantity/units) is a pure number.
  • Dimensional consistency: Both sides of an equation, all terms in a sum, and arguments of logarithms/exponentials must have matching or dimensionless character.

🧮 Fundamental concept

🧮 Physical quantity = numerical value × units

Physical quantity = numerical value × units

  • A physical quantity (unless dimensionless) is the product of a pure number and one or more units.
  • The symbol for a quantity does not lock you into particular units.
  • Example: Density ρ can be expressed as 9.970 × 10² kg m⁻³ or as 0.9970 g cm⁻³; both describe the same physical property.

🔢 Ratio of quantity to units is a pure number

  • Dividing both sides of ρ = 0.9970 g cm⁻³ by 1 g cm⁻³ gives ρ/(g cm⁻³) = 0.9970.
  • This ratio is "the number of grams per cubic centimeter."
  • In general: (quantity / units) = pure number representing how many of those units.
  • Don't confuse: Calling ρ itself "the number of grams per cubic centimeter" is incorrect because that phrase refers to a specific unit choice; only the ratio ρ/(g cm⁻³) is that number.
  • Similarly, n is not "the number of moles"; n/mol is the number of moles.

📊 Practical use in tables and graphs

  • Column headings and axis labels often show the ratio (quantity/units) so that entries are pure numbers.
  • Example: A table column headed ρ/(g cm⁻³) contains only numerical values like 0.9970, 1.234, etc.

⚙️ Algebraic manipulation and unit conversion

⚙️ Conversion factors equal to 1

  • From definitions like 1 J = 1 N·m and 1 Pa = 1 N m⁻², you can form ratios: (1 N·m / J) = 1 and (1 Pa / N m⁻²) = 1.
  • Multiplying a quantity by such a conversion factor changes its units but not its value.
  • Example: 3.099 × 10³ J m⁻³ × (1 N·m / J) × (1 Pa / N m⁻²) = 3.099 × 10³ Pa.

⚙️ SI coherence: all conversion factors are unity

  • SI base units and SI derived units (without prefixes) are coherent: values expressed in different combinations of these units have the same numerical value.
  • Practical implication: Enter numerical values in SI units into a calculator, and the result is automatically in SI units—no conversion factors needed.
  • Example: Using the ideal gas equation p = nRT/V with n = 5.000 mol, R = 8.3145 J K⁻¹ mol⁻¹, T = 298.15 K, V = 4.000 m³ gives p = 3.099 × 10³ J m⁻³, which is immediately 3.099 × 10³ Pa without extra conversion steps.

🔄 Converting to non-SI units

  • When you need a non-SI unit, apply a conversion factor from the definition.
  • Example: To convert 3.099 × 10³ Pa to torr, use (760 Torr / 101,325 Pa) = 1, yielding 23.24 Torr.

🧪 Dimensional analysis and consistency checks

🧪 Rules for dimensional consistency

  1. Both sides of an equation must have the same dimensions.
  2. All terms in a sum or difference must have the same dimensions.
  3. Logarithms, exponentials, and their arguments must be dimensionless.
  4. Exponents must be dimensionless.

🧪 Differentials and derivatives

  • A differential (e.g., df) is an infinitesimal quantity; if one side of an equation is infinitesimal, the other must be too.
  • Valid: df = a dx + b dy (both sides infinitesimal).
  • Invalid: df = ax + b dy (mixes finite and infinitesimal).
  • Dimensions of derivatives and integrals:
    • dp/dT and (∂p/∂T)V have dimensions of p/T.
    • (∂²p/∂T²)V has dimensions of p/T².
    • T dT has dimensions of T².

🧪 Example applications

ScenarioExplanationDimensional outcome
Gas constant R in J K⁻¹ mol⁻¹R has dimensions energy/(temperature × amount)RT has dimensions energy/amount; nRT has dimensions energy
Expression nRT ln(p/p°)Logarithm must be dimensionlessp° must have dimensions of pressure so p/p° is dimensionless; whole expression has dimensions of energy
Constants in van der Waals equationEach term must have dimensions of pressureAnalyze each term to deduce dimensions of a and b
  • Don't confuse: The symbol R or n does not carry units; only when you write R = 8.3145 J K⁻¹ mol⁻¹ do you specify units for a calculation.

📐 Prefixes and unit construction

📐 SI prefixes

  • Any SI unit symbol (except kg and °C) may be preceded by a prefix (Table 1.4: milli-, centi-, kilo-, mega-, etc.).
  • The prefix + unit symbol forms a new symbol that can be raised to a power without parentheses.
  • Examples:
    • 1 mg = 1 × 10⁻³ g
    • 1 cm = 1 × 10⁻² m
    • 1 cm³ = (1 × 10⁻² m)³ = 1 × 10⁻⁶ m³

📐 Practical note

  • The gram symbol g may take a prefix (e.g., mg, kg), but kg itself cannot (the kilogram is the base unit).
  • This construction simplifies expressing very large or very small quantities while maintaining dimensional clarity.
3

Dimensional Analysis

1.3 Dimensional Analysis

🧭 Overview

🧠 One-sentence thesis

Dimensional analysis provides a systematic method to check equations for errors and determine the dimensions of unknown constants by ensuring dimensional consistency throughout all mathematical operations.

📌 Key points (3–5)

  • Core principle: Both sides of an equation and all terms in a sum or difference must have the same dimensions.
  • Special rules: Logarithms, exponentials, their arguments, and quantities used as powers must be dimensionless.
  • Derivatives and integrals have dimensions: The derivative dp/dT has dimensions of p/T; the integral of T dT has dimensions of T².
  • Common confusion: Infinitesimal quantities (differentials like df) must match on both sides—mixing finite and infinitesimal terms (e.g., df = ax + bdy) is invalid.
  • Practical use: Dimensional analysis reveals the dimensions of unknown constants in equations and catches dimensional errors before calculation.

🔍 Fundamental rules of dimensional consistency

⚖️ Equation sides must match

Both sides of an equation have the same dimensions.

  • This is the most basic check: if the left side has dimensions of pressure, the right side must also have dimensions of pressure.
  • Example: In p = nRT/(V - nb) - n²a/V², both the left side (pressure) and the entire right side must have pressure dimensions.

➕ Terms in sums and differences must match

All terms of a sum or difference have the same dimensions.

  • When you add or subtract quantities, each term must have identical dimensions.
  • Example: In V - nb, the term nb must have dimensions of volume because V has dimensions of volume.
  • This rule helps identify the dimensions of unknown constants.

🔢 Dimensionless requirements

Logarithms and exponentials, and arguments of logarithms and exponentials, are dimensionless. A quantity used as a power is dimensionless.

  • ln(x), exp(x), and x in these expressions must all be pure numbers without units.
  • Example: In nRT ln(p/p°), the ratio p/p° must be dimensionless, which means p° must have dimensions of pressure to cancel the dimensions of p.
  • Powers like T^(1/2) are allowed, but the exponent itself (1/2) is dimensionless.

📐 Dimensions of calculus operations

📉 Derivatives and partial derivatives

  • Single derivative: dp/dT and (∂p/∂T)_V both have the same dimensions as p/T.
  • Second derivative: (∂²p/∂T²)_V has dimensions of p/T².
  • The derivative "divides" the dimensions of the numerator by the dimensions of the denominator.

∫ Integrals

  • The integral ∫T dT has the same dimensions as T².
  • Integration "multiplies" the dimensions of the integrand by the dimensions of the differential.

⚠️ Infinitesimal consistency

  • If one side of an equation is an infinitesimal quantity (a differential), the other side must also be infinitesimal.
  • Valid: df = a dx + b dy (where ax and by have the same dimensions as f)—all terms are infinitesimal.
  • Invalid: df = ax + b dy—mixing an infinitesimal (df, bdy) with a finite term (ax) makes no mathematical sense.

🧪 Worked examples from thermodynamics

🧪 Example 1: Gas constant R and energy

  • The gas constant R has units of J K⁻¹ mol⁻¹.
  • Dimensions: energy / (thermodynamic temperature × amount).
  • RT: dimensions of energy/amount.
  • nRT: dimensions of energy (because n has dimensions of amount, which cancels the amount in the denominator of RT).
  • These products appear frequently in thermodynamic expressions.

🧪 Example 2: Logarithmic expression nRT ln(p/p°)

Question: What are the dimensions of nRT ln(p/p°) and of p° in this expression?

  • The entire quantity has the same dimensions as nRT, which is energy.
  • Why? Because the logarithm ln(p/p°) is dimensionless.
  • For the logarithm to be dimensionless, its argument p/p° must be dimensionless.
  • Therefore, p° must have dimensions of pressure to cancel the dimensions of p.

🧪 Example 3: Van der Waals equation constants

Equation: p = nRT/(V - nb) - n²a/V²

Finding dimensions of b:

  • The term nb is subtracted from V, so nb must have dimensions of volume.
  • Therefore, b has dimensions of volume/amount.

Finding dimensions of a:

  • The right side is a difference of two terms, so both terms must have the same dimensions as the left side (pressure).
  • The second term n²a/V² has dimensions of pressure.
  • Rearranging: a has dimensions of pressure × volume² / amount².

🧪 Example 4: Partial derivative equation

Equation: (∂ ln x / ∂T)_p = y/R

Question: What are the SI units of y?

  • ln x is dimensionless, so the left side has dimensions of 1/T.
  • SI units of the left side: K⁻¹.
  • The right side must also have units of K⁻¹.
  • R has units J K⁻¹ mol⁻¹.
  • Therefore, y must have SI units of J K⁻² mol⁻¹ to make y/R have units of K⁻¹.

🎯 Practical application strategy

🎯 How to find dimensions of unknown constants

  1. Identify terms that are added or subtracted—they must have matching dimensions.
  2. Identify arguments of logarithms or exponentials—they must be dimensionless.
  3. Use the requirement that both sides of the equation have the same dimensions.
  4. Solve algebraically for the dimensions of the unknown constant.

🎯 Catching errors

  • Check every equation for dimensional consistency before performing calculations.
  • Verify that sums and differences involve terms with identical dimensions.
  • Ensure logarithms and exponentials have dimensionless arguments.
  • Confirm that infinitesimal and finite quantities are not mixed inappropriately.
4

2.1 The System, Surroundings, and Boundary

2.1 The System, Surroundings, and Boundary

🧭 Overview

🧠 One-sentence thesis

Thermodynamics requires us to precisely define a system—any chosen three-dimensional region of space—along with its boundary and surroundings, because these definitions determine what matter and energy can cross into or out of the system and thus control how we analyze its behavior.

📌 Key points (3–5)

  • What a system is: any three-dimensional region of physical space we choose to focus on; the rest of the universe is the surroundings, and the boundary separates them.
  • The boundary is flexible: it may coincide with real physical surfaces (walls, phase interfaces) or be an imaginary surface; size and shape can change over time.
  • Open vs closed systems: if matter crosses the boundary in either direction, the system is open; otherwise it is closed.
  • Diathermal vs adiabatic boundaries: diathermal boundaries allow heat transfer; adiabatic boundaries do not (perfect insulation).
  • Common confusion: an isolated system exchanges no matter, heat, or work with surroundings, but it may still experience forces or external fields (e.g., gravity) from the surroundings.

🔲 Defining the system, surroundings, and boundary

🔲 What a system is

A thermodynamic system is any three-dimensional region of physical space on which we wish to focus our attention.

  • Classical thermodynamics treats matter as a continuous medium rather than discrete particles, focusing on macroscopic properties of large aggregates of molecules, atoms, and ions.
  • The surroundings are the rest of the physical universe outside the system.
  • The boundary is the closed three-dimensional surface that encloses the system and separates it from the surroundings.

🎯 The boundary is arbitrary but must be precise

  • The boundary may coincide with real physical surfaces:
    • The interface between two phases.
    • The inner or outer surface of a flask or vessel wall.
  • Alternatively, part or all of the boundary may be an imaginary intangible surface in space, unrelated to any physical structure.
  • The size and shape of the system (defined by its boundary) may change in time.
  • Key point: Our choice of the system is arbitrary, but we must know exactly what that choice is.

👤 Experimenters are part of the surroundings

  • We (the experimenters) are part of the surroundings, not the system.
  • We influence the system indirectly through its interaction with the surroundings.
  • We manipulate the surroundings directly with physical devices under our control.

🧩 Subsystems and supersystems

  • A system may be divided into subsystems for some purposes.
  • Two or more systems may be combined and treated as a supersystem.

🚪 Types of systems: open, closed, and isolated

🚪 Open vs closed systems

If over the course of time matter is transferred in either direction across the boundary, the system is open; otherwise it is closed.

  • Open system: matter crosses the boundary.
    • Matter may pass through a stationary boundary, or
    • The boundary may move through matter that is fixed in space.
  • Closed system: no matter crosses the boundary.

🔥 Diathermal vs adiabatic boundaries

If the boundary allows heat transfer between the system and surroundings, the boundary is diathermal.

An adiabatic boundary is a boundary that does not allow heat transfer.

  • An adiabatic boundary can be ensured (in principle) by surrounding the system with an adiabatic wall:
    • Perfect thermal insulation.
    • Perfect radiation shield.
  • Example: A gas in a thermally-insulated container has an adiabatic boundary; a gas in a metal container has a diathermal boundary.

🔒 Isolated systems

An isolated system is one that exchanges no matter, heat, or work with the surroundings, so that the system's mass and total energy remain constant over time.

  • A closed system with an adiabatic boundary, constrained to do no work and to have no work done on it, is an isolated system.
  • Don't confuse: An isolated system may still interact with the surroundings in the sense that:
    • Forces exist between the system and surroundings (e.g., a gas exerts force on rigid walls, and the walls exert equal and opposite force on the gas).
    • The system may experience a constant external field, such as a gravitational field.
  • These interactions do not violate the "isolated" condition because no matter, heat, or work is exchanged.
System typeMatter exchange?Heat exchange?Work exchange?
OpenYesMay or may notMay or may not
ClosedNoMay or may notMay or may not
IsolatedNoNoNo

📦 The concept of a body

📦 What a body is

The term body usually implies a system, or part of a system, whose mass and chemical composition are constant over time.

  • A body is a special case: it does not exchange matter with surroundings (closed) and its composition does not change internally.
  • Example: A sealed container of gas with fixed composition is a body; an open beaker where water evaporates is not.
5

Phases and Physical States of Matter

2.2 Phases and Physical States of Matter

🧭 Overview

🧠 One-sentence thesis

Phases are distinguished by whether intensive properties vary continuously or discontinuously, and physical states (solid vs. fluid) are rigorously classified by how matter responds to shear stress rather than by vague descriptions of rigidity or compressibility.

📌 Key points (3–5)

  • What a phase is: a region where intensive properties are either uniform throughout or vary continuously, meeting other phases at interface surfaces where properties change discontinuously.
  • How to classify physical states: the primary distinction is solid vs. fluid based on response to shear stress; fluids are then further classified as liquid, gas, or supercritical fluid.
  • Common confusion: traditional labels (solid = rigid, liquid = incompressible, gas = compressible) differ only in degree and fail for intermediate cases; the rigorous approach uses shear stress response.
  • Nonuniform regions: a region with continuously varying properties can be treated either as a single nonuniform phase or as infinitely many infinitesimal uniform phases.

🔬 What defines a phase

🔬 Phase definition and uniformity

A phase is a region of the system in which each intensive property (such as temperature and pressure) has at each instant either the same value throughout (a uniform or homogeneous phase), or else a value that varies continuously from one point to another.

  • Unless stated otherwise, "phase" in this context means a uniform phase.
  • Intensive properties have the same value everywhere in a uniform phase.
  • Two different phases meet at an interface surface, where intensive properties have a discontinuity or change value over a small distance.

🌐 Isotropic vs anisotropic phases

  • Some intensive properties (e.g., refractive index, polarizability) can have directional characteristics.
  • A uniform phase may be:
    • Isotropic: same property values in all directions.
    • Anisotropic: directional variation (e.g., some solids and liquid crystals).
  • A vacuum is a uniform phase of zero density.

📐 Handling nonuniform regions

  • Example: a tall column of gas in a gravitational field where density decreases with altitude.
  • Two treatment options:
    1. Treat as a single nonuniform phase.
    2. Treat as an infinite number of uniform phases, each infinitesimal in one or more dimensions.

🧱 Classifying physical states

🧱 The inadequacy of traditional descriptions

  • Common informal labels:
    • Solid: relatively rigid.
    • Liquid: easily deformed and relatively incompressible.
    • Gas: easily deformed and easily compressed.
  • Problem: these descriptions differ only in degree and cannot classify intermediate cases.
  • Don't confuse: rigidity and compressibility are continuous properties, not sharp boundaries.

⚙️ Shear stress as the rigorous criterion

Shear stress: a tangential force per unit area exerted on matter on one side of an interior plane by the matter on the other side.

  • Shear stress is produced by applying tangential forces to parallel surfaces of a phase.
  • The primary distinction is between solid and fluid, based on response to shear stress.
  • Additional criteria then classify fluids as liquid, gas, or supercritical fluid.

🔧 How solids and fluids respond to shear stress

🔧 Solid response: deformation

  • A solid responds to shear stress by undergoing momentary relative motion of its parts, resulting in deformation (a change of shape).
  • If the applied shear stress is constant and small (not large enough to cause creep or fracture):
    • The solid quickly reaches a certain degree of deformation.
    • The degree of deformation depends on the magnitude of the stress.
    • The solid maintains this deformation without further change as long as the shear stress continues.
  • Microscopic mechanism: deformation requires relative movement of adjacent layers of particles (atoms, molecules, or ions).
  • Example: applying tangential forces to opposite surfaces of a block causes it to deform into a parallelogram shape, which it holds as long as the force is applied.

🌊 Fluid response (implied)

  • The excerpt defines a fluid implicitly as a phase that does not behave like a solid under shear stress.
  • Fluids are further classified as liquid, gas, or supercritical fluid using additional criteria (not detailed in this excerpt).
  • Don't confuse: the solid/fluid distinction is based on whether the phase reaches a stable deformation under constant small shear stress, not on how "hard" or "soft" it feels.
6

Some Basic Properties and Their Measurement

2.3 Some Basic Properties and Their Measurement

🧭 Overview

🧠 One-sentence thesis

Mass, amount of substance, volume, density, pressure, and temperature are fundamental macroscopic properties that can be measured with various instruments, each requiring careful calibration and accounting for systematic effects like buoyancy or gravitational field variations.

📌 Key points (3–5)

  • Mass measurement: Modern balances measure gravitational force on a body; the most accurate measurements account for air buoyancy and trace calibration to national standards.
  • Amount of substance (moles): Although the mole is a counting unit, chemists measure it by weighing, using the relationship between mass and relative atomic/molecular mass.
  • Molar mass: Defined as mass divided by amount (M = m/n); it connects the extensive property (mass) to the intensive property (amount).
  • Common confusion: The 2019 SI revision redefined the kilogram using the Planck constant instead of a physical prototype, but for most purposes the old definition (twelve grams of carbon-12 = one mole) remains valid.
  • Measurement methods: Each property has multiple measurement techniques with different typical uncertainties, from analytical balances (±0.1 mg) to vibrating-tube densimeters (±0.01 mg/mL).

⚖️ Mass and its measurement

⚖️ What mass measurement involves

  • The SI unit of mass is the kilogram.
  • Practical measurement uses a balance that exploits the downward gravitational force exerted by Earth on the body.
  • Two main types:
    • Classic balance: beam and knife-edge arrangement that compares gravitational force on the unknown body with force on a known mass.
    • Modern balance (scale): uses a strain gauge or similar device to directly measure gravitational force; must be calibrated with known masses.

🎈 Accounting for buoyancy

  • The most accurate measurements take into account the buoyancy of the body and calibration masses in air.
  • Buoyancy is the upward force exerted by the surrounding air, which reduces the apparent weight.
  • Ignoring buoyancy introduces systematic error, especially for low-density objects.

🔬 The Kibble balance and the new kilogram definition

  • The 2019 SI revision replaced the international prototype kilogram with a definition based on the Planck constant h.
  • The Kibble balance (formerly watt balance) is the method of choice for applying this definition, achieving uncertainty of several parts in 10⁸.
  • How it works:
    1. Balance step: A test weight of unknown mass m is placed on one side; a current I flows through a coil in a magnetic field to generate an upward electromagnetic force. The balance condition is: gravitational force = electromagnetic force, or mg = BlI (where B = magnetic flux density, l = coil wire length, g = acceleration of free fall).
    2. Calibration step: Remove the test weight, turn off the current, and move the coil vertically at constant speed v through the magnetic field. This induces a potential difference Φ = Blv.
    3. Calculation: Eliminate the product Bl from the two equations to get m = IΦ / gv. The current I and potential Φ are measured with high precision using methods that depend on the defined value of h; g is measured with a gravimeter.
  • Example: The international prototype's mass is found to be 1 kg to within 1 × 10⁻⁸ kg using this method.

📊 Typical measurement uncertainties

InstrumentTypical valueApproximate uncertainty
Analytical balance100 g0.1 mg
Microbalance20 mg0.1 µg

🧪 Amount of substance (moles)

🧪 What the mole measures

  • The SI unit of amount of substance is the mole.
  • Although the mole is a counting unit, chemists measure it by weighing, not by counting individual atoms or molecules.

🔢 The carbon-12 reference

The previous definition of the mole (still valid for most purposes): twelve grams of carbon-12 contains one mole of atoms.

  • Relative atomic mass (atomic weight) Aᵣ: dimensionless quantity equal to the atomic mass relative to Aᵣ = 12 for carbon-12.
  • Relative molecular mass (molecular weight) Mᵣ: dimensionless quantity equal to the molecular mass relative to carbon-12.

🧮 Calculating amount from mass

The amount n of a substance of mass m can be calculated from:

  • For atoms: n = m / (Aᵣ g/mol)
  • For molecules: n = m / (Mᵣ g/mol)

Example: If you have 24 grams of carbon-12, the amount is n = 24 g / (12 g/mol) = 2 mol.

📐 Molar mass

Molar mass M: defined as the mass divided by the amount, M = m / n.

  • Molar mass is an intensive property (does not depend on the size of the sample).
  • It connects the extensive property (mass) to the amount.
  • The symbol M for molar mass is noted as an exception (the excerpt cuts off here, but likely refers to notation conventions).

⚠️ Don't confuse: 2019 SI revision impact

  • The 2019 SI revision redefined the kilogram and made a negligible change to calculations involving the mole.
  • For most purposes, the old definition (twelve grams of carbon-12 = one mole) remains valid.
  • The new definition uses the Avogadro constant as a defined value, but practical calculations are essentially unchanged.

📏 Other basic properties and measurement methods

📏 Volume measurement

InstrumentTypical valueApproximate uncertainty
Pipet, Class A10 mL0.02 mL
Volumetric flask, Class A1 L0.3 mL

🧊 Density measurement

InstrumentTypical valueApproximate uncertainty
Pycnometer, 25-mL capacity1 g/mL2 mg/mL
Magnetic float densimeter1 g/mL0.1 mg/mL
Vibrating-tube densimeter1 g/mL0.01 mg/mL
  • Density is an intensive property (mass per unit volume).
  • Different methods offer different precision; vibrating-tube densimeters are the most precise listed.

🌡️ Pressure measurement

InstrumentTypical valueApproximate uncertainty
Mercury manometer or barometer760 Torr0.001 Torr
Diaphragm gauge100 Torr1 Torr

🌡️ Temperature measurement

InstrumentTypical valueApproximate uncertainty
Constant-volume gas thermometer10 K0.001 K
Mercury-in-glass thermometer300 K0.01 K
Platinum resistance thermometer300 K0.0001 K
Monochromatic optical pyrometer1300 K0.03 K
  • Different temperature ranges and applications require different instruments.
  • Platinum resistance thermometers offer the highest precision at room temperature.
  • Optical pyrometers are used for high-temperature measurements (e.g., 1300 K).

🔗 Calibration and traceability

🔗 National and international standards

  • The accuracy of calibration masses should be traceable to a national standard kilogram.
  • In the United States, the national standard is maintained at NIST (National Institute of Standards and Technology) in Gaithersburg, Maryland.
  • National standards are ultimately traceable to the international prototype (before 2019) or the new definition based on fundamental constants (after 2019).

🔗 Why traceability matters

  • Traceability ensures that measurements made in different laboratories or at different times can be compared reliably.
  • It provides a chain of calibrations linking everyday measurements to the most fundamental definitions.
  • Example: An analytical balance in a lab is calibrated with weights that were calibrated against NIST standards, which in turn are calibrated using a Kibble balance that implements the SI definition of the kilogram.

Budget: budget:token_budget1000000</budget:token_budget> Tokens used: 7022 Remaining: 992978

7

The State of the System

2.4 The State of the System

🧭 Overview

🧠 One-sentence thesis

Precise measurement of fundamental physical properties—mass, amount of substance, volume, density, pressure, and temperature—requires specialized apparatus and carefully defined scales, with temperature being unique in having multiple valid definitions and scales rather than a single operational method.

📌 Key points (3–5)

  • Mass measurement at high precision uses elaborate apparatus (Kibble balance) that relates mechanical force to electromagnetic force via the Planck constant.
  • Amount of substance (moles) is measured by weighing, not counting, using relative atomic/molecular mass and the carbon-12 reference.
  • Temperature is fundamentally different: unlike other properties, it has no single unique definition and requires choosing a temperature scale with an operational measurement method.
  • Common confusion: temperature scales vs. temperature units—thermodynamic temperature and ideal-gas temperature are proportional and become identical when using the same unit (kelvin), but the Celsius scale is the thermodynamic scale shifted by exactly 273.15 K.
  • Fixed-point systems (triple points, melting points) provide the most reproducible reference temperatures for calibration.

⚖️ Mass measurement

⚖️ The Kibble balance method

A Kibble balance (formerly watt balance) is an elaborate apparatus that measures mass by balancing gravitational force against electromagnetic force.

  • Why it matters: achieves uncertainty of several parts in 10⁸, allowing the international prototype kilogram to be measured as 1 kg within 1 × 10⁻⁸ kg.
  • How it works (two-step process):
    1. Balance step: test weight of unknown mass m is added; current I through a coil in a magnetic field is adjusted until gravitational force equals electromagnetic force: mg = BlI (where g = acceleration of free fall, B = magnetic flux density, l = wire length).
    2. Calibration step: test weight removed, current off, coil moved vertically at constant speed v through the field, inducing potential difference Ʌ = Blv.
  • Final calculation: eliminating the product Bl from both equations gives m = IƁ/gv.
  • Key insight: I and Ʌ are measured to very high precision using Josephson and quantum Hall effects, which require the defined value of Planck constant h; g is measured with a gravimeter.

📊 Typical measurement methods

InstrumentTypical valueApproximate uncertainty
Analytical balance100 g0.1 mg
Microbalance20 mg0.1 μg

🧪 Amount of substance and molar quantities

🧪 The mole and relative mass

The mole is a counting unit, but an amount in moles is measured by weighing, not counting.

  • Reference standard: twelve grams of carbon-12 contains one mole of atoms (valid for most purposes after 2019 SI revision, which makes negligible change to calculations).
  • Relative atomic mass (atomic weight) Aᵣ: dimensionless quantity equal to atomic mass relative to Aᵣ = 12 for carbon-12.
  • Relative molecular mass (molecular weight) Mᵣ: dimensionless quantity for molecular substances, molecular mass relative to carbon-12.

🧮 Calculating amount from mass

Amount n of a substance of mass m is calculated from:

  • n = m / (Aᵣ g mol⁻¹) for atoms
  • n = m / (Mᵣ g mol⁻¹) for molecules

🔢 Molar mass

Molar mass M is defined as mass divided by amount: M = m/n.

  • Key relationship: the numerical value of molar mass in g mol⁻¹ equals the relative atomic or molecular mass: M/(g mol⁻¹) = Aᵣ or M/(g mol⁻¹) = Mᵣ.
  • Example: if a substance has Mᵣ = 18, its molar mass is 18 g mol⁻¹.

📏 Volume and density

📏 Volume measurement

  • Common methods: burets, pipets, volumetric flasks (Class A glassware has NIST specifications).
  • SI unit: cubic meter (m³), but chemists commonly use liters (L) and milliliters (mL).
  • Liter definition: one cubic decimeter (dm³); 1 m³ = 10³ L = 10⁶ mL; 1 mL = 1 cm³.
  • Historical note: before 1964, the liter was defined as the volume of 1 kg of water at 3.98 °C (maximum density), making 1 L = 1.000028 dm³—numerical values from before 1964 may need small corrections.
InstrumentTypical valueApproximate uncertainty
Pipet, Class A10 mL0.02 mL
Volumetric flask, Class A1 L0.3 mL

📐 Density

Density is defined as the ratio of mass to volume: ρ = m/V.

  • Intensive property: does not depend on system size.
  • Relation to molar volume: Vₘ = M/ρ (molar volume inversely proportional to density).
  • Measurement methods (three examples illustrated):
    1. Pycnometer: glass vessel with capillary stopper; filled, brought to temperature, dried, and weighed.
    2. Magnetic float densimeter: buoy containing magnet is held in position by solenoid; required current depends on liquid density.
    3. Vibrating-tube densimeter: liquid-filled metal U-tube vibrates; resonance frequency is a function of liquid mass in tube.
InstrumentTypical valueApproximate uncertainty
Pycnometer, 25-mL1 g mL⁻¹2 mg mL⁻¹
Magnetic float densimeter1 g mL⁻¹0.1 mg mL⁻¹
Vibrating-tube densimeter1 g mL⁻¹0.01 mg mL⁻¹

🌡️ Pressure

🌡️ Definition and characteristics

Pressure is the normal component of stress exerted by an isotropic fluid on a surface element.

  • The surface can be an interface between fluid and another phase, or an imaginary dividing plane within the fluid.
  • Usually positive, but negative pressure is possible in liquids under tension (e.g., mercury column suspended in capillary tube with no vapor bubble)—this is unstable and can cause spontaneous vaporization.
  • Don't confuse: pressure in a fluid vs. stress in a solid—within a solid, forces are described by nine components of a stress tensor; saying a solid "is at" a certain pressure means that is the hydrostatic pressure on its exterior surface.

📊 Units of pressure

  • SI unit: pascal (Pa) = 1 newton per square meter (N/m²).
  • Non-SI units commonly used by chemists:
    • Millimeter of mercury (mmHg): pressure exerted by 1 mm column of fluid with density exactly 13.5951 g cm⁻³ (mercury at 0 °C) where g has standard value gₙ.
    • Torr: defined so that 1 atm = exactly 760 Torr.
    • Atmosphere (atm): exactly 1.01325 × 10⁵ Pa.
    • Bar: exactly 10⁵ Pa (approximately 1% smaller than 1 atm).
  • Practical identity: mmHg and Torr differ by less than 2 × 10⁻⁷ Torr.
  • Standard pressure p°: IUPAC recommends 1 bar (formerly 1 atm was used).
InstrumentTypical valueApproximate uncertainty
Mercury manometer/barometer760 Torr0.001 Torr
Diaphragm gauge100 Torr1 Torr

🌡️ Temperature: fundamentals and scales

🌡️ Why temperature is unique

Temperature does not have a single unique definition—unlike mass, volume, or pressure.

  • Any chosen definition requires a temperature scale described by an operational measurement method.
  • Only requirement: values should increase monotonically with physiological "hotness."
  • Example: a simple liquid-in-glass thermometer with equally spaced marks defines a temperature scale by linear interpolation between marks.
  • Measurement principle: placing thermometer and body in thermal contact may affect the body's temperature; the measured temperature is that of the body after thermal equilibrium is achieved.

🔄 Comparing temperatures and the zeroth law

  • Thermometer principle: temperatures of different bodies may be compared with a thermometer.
  • Key insight: if two bodies give the same thermometer reading (on any scale), they have the same temperature.
  • Significance: bodies with the same temperature, when placed in thermal contact, will be in thermal equilibrium (no changes in properties).

Zeroth law of thermodynamics (Maxwell, 1872): "Bodies whose temperatures are equal to that of the same body have themselves equal temperatures."

❄️ Fixed-temperature equilibrium systems

Triple points provide the most reproducible temperatures—both temperature and pressure have definite fixed values when solid, liquid, and gas phases of a pure substance coexist.

System typeReproducibilityNotes
Ice point (ice + air-saturated water at 1 atm)LowAffected by air bubbles in ice and varying air concentrations
Steam point (liquid + gas H₂O at 1 atm)LowSensitive to experimental pressure
Melting point (solid + liquid at controlled pressure)HigherMore reproducible
Triple point (solid + liquid + gas)HighestReproducibility within 10⁻⁴ K for water
  • Water triple-point cell: contains pure water (same isotopic composition as ocean water); air pumped out, sealed; thin ice layer formed and partially melted to create pure water film; thermometer bulb placed in inner well with ice water for thermal contact.

🌡️ Six temperature scales

🌡️ Ideal-gas and thermodynamic scales

Ideal-gas temperature scale: defined by gas thermometry measurements.

Thermodynamic temperature scale: defined by behavior of a theoretical Carnot engine.

  • Key relationship: the two scales are proportional to one another; values become identical when using the same temperature unit.
  • The kelvin (pre-2019 definition): specified that the triple point of water has thermodynamic temperature of exactly 273.16 kelvins; ideal-gas temperature set equal to the same value.
  • The kelvin (2019 SI revision): triple point temperature of water is determined experimentally by primary thermometry, resulting in 273.16 K within 1 × 10⁻⁷ K—no practical difference from old definition.
  • Convention: symbol T refers to thermodynamic temperature, but is used for both thermodynamic and ideal-gas temperature since they have identical values.

🌡️ Centigrade and Celsius scales

Obsolete centigrade scale: exactly 0 °C at ice point, exactly 100 °C at steam point, linear function of ideal-gas temperature.

Celsius scale: thermodynamic (or ideal-gas) temperature scale shifted by exactly 273.15 kelvins.

  • Temperature unit: degree Celsius (°C), identical in size to the kelvin.
  • Conversion formula: t/°C = T/K − 273.15
  • Reference points on Celsius scale:
    • Triple point of water: exactly 0.01 °C
    • Ice point: 0 °C within 0.0001 °C
    • Steam point: 99.97 °C

🌡️ International Temperature Scale of 1990 (ITS-90)

ITS-90 defines international temperature T₉₀, intended to be very close to thermodynamic temperature T.

  • Temperature range: 0.65 K to at least 1358 K.
  • Measurement methods (depending on range):
    • Vapor-pressure thermometry (0.65–5.0 K): formulas for T₉₀ in terms of vapor pressure of ³He and ⁴He.
    • Gas thermometry (3.0–24.5561 K)
    • Platinum-resistance thermometry (13.8033–1234.93 K)
    • Optical pyrometry (above 1234.93 K)
  • Calibration: assigns values to fourteen fixed temperatures (see table below) and provides interpolating functions for intermediate temperatures.
  • Status: temperatures defined by ITS-90 are exact with respect to the scale—values remain unchanged during the life of the scale.

📋 ITS-90 fixed temperatures

T₉₀ / KEquilibrium system
13.8033H₂ triple point
24.5561Ne triple point
54.3584O₂ triple point
83.8058Ar triple point
234.3156Hg triple point
273.16H₂O triple point
302.9146Ga melting point at 1 atm
429.7485In melting point at 1 atm
505.078Sn melting point at 1 atm
692.677Zn melting point at 1 atm
933.473Al melting point at 1 atm
1234.93Ag melting point at 1 atm
1337.33Au melting point at 1 atm
1357.77Cu melting point at 1 atm

🌡️ Provisional Low Temperature Scale of 2000 (PLST-2000)

  • Temperature range: 0.0009 K to 1 K.
  • Basis: melting temperature of solid ³He as a function of pressure (30–40 bar at these temperatures).
  • Status: temperatures are exact with respect to the scale.

🔬 Primary thermometry

🔬 What is primary thermometry

Primary thermometry: measurement of temperature based on fundamental physical principles.

  • Historical method: gas thermometry (until about 1960).
  • Modern methods: more accurate but require elaborate equipment, not convenient for routine measurements.
  • Required constants: Boltzmann constant k or gas constant R = Nₐk (where Nₐ is Avogadro constant); these are defining constants in 2019 SI revision, with fixed values ensuring consistency with the kelvin definition.

🔬 Gas thermometry method

Basis: ideal gas equation T = pV/nR.

  • Thermometric gas: usually helium (minimal deviations from ideal-gas behavior).
  • Constant-volume gas thermometer: bulb containing gas + means of measuring pressure.
    • Simple version: mercury manometer.
    • Sophisticated version: diaphragm pressure transducer between gas bulb and pressure measurement system.

🔬 Measurement procedure

Two-step process to determine unknown temperature:

  1. Reference measurement: gas brought into thermal equilibrium with reference system of known temperature T₁ (e.g., from ITS-90 table); pressure p₁ measured.
  2. Unknown measurement: gas brought into thermal equilibrium with system whose temperature T₂ is to be measured; pressure p₂ measured.
  • If gas were ideal: nR = p₁V₁/T₁ = p₂V₂/T₂, giving T₂ = T₁(p₂V₂/p₁V₁).
  • In practice: gas approaches ideal behavior only at low pressure, so a series of paired measurements is made, changing the amount of gas before each pair (measurements extrapolated to zero pressure).
  • Volume correction: gas volume is nearly the same in both measurements except for small changes due to effects of T and p on gas bulb dimensions.
InstrumentTypical valueApproximate uncertainty
Constant-volume gas thermometer10 K0.001 K
Mercury-in-glass thermometer300 K0.01 K
Platinum resistance thermometer300 K0.0001 K
Monochromatic optical pyrometer1300 K0.03 K
8

Processes and Paths

2.5 Processes and Paths

🧭 Overview

🧠 One-sentence thesis

The thermodynamic state of a system is defined by the values of a minimum set of independent state functions at a given instant, and all other properties become dependent variables determined by those independent variables.

📌 Key points (3–5)

  • State functions: properties whose values depend only on the current state, not on the system's history.
  • Independent vs dependent variables: a minimum number of state functions serve as independent variables; all other state functions become dependent variables with unique, reproducible values.
  • State vs physical state: the thermodynamic "state" is not the same as "physical state" (solid/liquid/gas); a change of state means any property change, not necessarily a phase transition.
  • Common confusion: don't confuse the thermodynamic state (defined by property values) with the phase or state of aggregation of matter.
  • Number of independent variables: for a given system under specified conditions, the number of independent variables is fixed (e.g., four for a two-component single-phase system).

🎯 What is the thermodynamic state?

🎯 Definition and core idea

The thermodynamic state of the system: the condition described by the values of relevant macroscopic properties at a given instant.

  • At each instant, the system is in some definite state.
  • We describe the state using values of macroscopic properties we consider relevant.
  • Whenever any relevant property value changes, the state has changed.
  • If all relevant properties return to their previous values, the system has returned to its previous state.

⚠️ Don't confuse state with physical state

  • Thermodynamic state: defined by property values (temperature, pressure, composition, etc.).
  • Physical state (state of aggregation): whether a phase is solid, liquid, or gas.
  • A "change of state" means a change in the thermodynamic state—not necessarily a phase transition.

🔧 State functions

🔧 What state functions are

State functions (or state variables, state parameters): properties whose values at each instant depend only on the state of the system at that instant, not on past or future history.

  • Examples from the excerpt: temperature T, pressure p, amounts of substances (n_A, n_B), volume V, mass m, density ρ, mole fraction x_B, osmotic pressure Π, refractive index n_D.
  • Properties we consider irrelevant (e.g., shape of the system) are not state functions.

🔧 Why state functions matter

  • Once the state is defined, every state function has one definite, reproducible value.
  • State functions allow us to describe the system completely without needing to know how it reached that state.

🧮 Independent and dependent variables

🧮 How independent variables define the state

  • A certain minimum number of state functions are chosen as independent variables.
  • These independent variables must be consistent with:
    • The physical nature of the system.
    • Any conditions or constraints (e.g., closed system → constant mass; rigid boundary → constant volume).
    • Equations of state (if applicable).

🧮 Dependent variables follow automatically

  • Once independent variables are set to particular values, every other state function becomes a dependent variable.
  • Each dependent variable can have only one definite, reproducible value for a given set of independent variable values.
  • Example: For a single-phase pure substance with T, p, and n as independent variables:
    • Volume V is determined by the equation of state.
    • Mass m = nM (where M is molar mass).
    • Molar volume V_m = V/n.
    • Density ρ = nM/V.

🧮 The number of independent variables is fixed

  • For a given system under specified conditions, the number of independent variables is fixed.
  • Example from the excerpt: an aqueous sucrose solution (two-component, single-phase) has four independent variables.
  • You cannot arbitrarily choose just any properties as independent variables—they must be consistent with the system's nature and constraints.

📊 Example: aqueous sucrose solution

📊 The system and its state

The excerpt provides a concrete example: an aqueous sucrose solution (water = A, sucrose = B) in a particular state.

PropertyValueRole
Temperature T293.15 KIndependent variable
Pressure p1.01 barIndependent variable
Amount of water n_A39.18 molIndependent variable
Amount of sucrose n_B1.375 molIndependent variable
Volume V1000 cm³Dependent (from equation of state)
Mass m1176.5 gDependent (calculated)
Density ρ1.1765 g/cm³Dependent (calculated)
Mole fraction of sucrose x_B0.03390Dependent (calculated)
Osmotic pressure Π58.2 barDependent (solution property)
Refractive index n_D1.400Dependent (optical property)

📊 How the variables relate

  • The first four properties (T, p, n_A, n_B) are chosen as independent variables.
  • These four values suffice to define the state for most purposes.
  • Experimental measurements confirm: whenever these four have these particular values, each other property has one definite value.
  • You cannot alter any dependent variable without changing at least one independent variable.

📊 Other possible choices

  • The independent variables can be chosen differently, e.g.:
    • T, p, V, and x_B.
    • p, V, ρ, and x_B.
  • However, the number of independent variables (four for this system) remains the same.
  • Not every set of four properties can serve as independent variables—the choice must be physically meaningful.

📊 Innumerable dependent variables

  • Beyond the six dependent variables listed in the table, the system has many more: energy, isothermal compressibility, heat capacity at constant pressure, etc.
  • All are determined once the independent variables are fixed.
9

The Energy of the System

2.6 The Energy of the System

🧭 Overview

🧠 One-sentence thesis

The internal energy U is defined as the energy of a system measured in a reference frame that makes U a state function depending only on the system's state, whereas the total system energy E_sys measured in a lab frame can include kinetic and potential energy contributions from the system's motion as a whole.

📌 Key points (3–5)

  • State vs path functions: State functions (like temperature, volume, energy) have values at one instant; path functions (like heat, work) refer to an interval of time and depend on the process path.
  • Energy depends on reference frame: The kinetic energy and total energy of a system depend on which coordinate system (reference frame) is used to measure velocities and positions.
  • Internal energy U vs system energy E_sys: U is measured in a "local frame" that moves with the system so U depends only on the system's state; E_sys in a lab frame includes motion and position of the system as a whole.
  • Common confusion: When a beaker of water slides or is lifted, E_sys changes but U does not, because the state (T and p) is unchanged—motion and height affect E_sys but not the internal state.
  • Absolute values are impractical: Only changes in internal energy (ΔU) are useful; calculating U from mass via Einstein's relation introduces enormous uncertainty, and mass changes in chemical processes are too small to detect.

🔄 State functions vs path functions

🔄 What distinguishes them

State function: a property whose value refers to one instant of time and depends only on the current state, not on how the system reached that state.

Path function: a quantity whose value refers to an interval of time and depends on the process path taken.

  • Examples from the excerpt:
    • State functions: temperature, volume, energy
    • Path functions: heat (q), work (w)
  • The integral of an infinitesimal state function gives a definite change; the integral of a path function gives a net quantity that depends on the path.

🏔️ Mountain analogy

  • Elevation is like a state function: at each instant the climber is at a definite elevation; the elevation change depends only on starting and ending points, not on which trail was used.
  • Distance traveled is like a path function: the total distance depends on which trail the climber took.
  • This analogy clarifies why state functions are path-independent and path functions are not.

🧮 Notation for infinitesimals

  • The excerpt uses ∂q and ∂w (with a bar through the "d") for infinitesimal quantities of path functions.
  • These are called inexact differentials because their integrals depend on the path.
  • Don't confuse: ordinary differentials (like dT, dV) are exact; ∂q and ∂w are inexact.
  • Alternative notations mentioned: dq and dw, Δq and Δw, δq and δw.

🎯 Reference frames and energy

🎯 What is a reference frame

Reference frame: a system of coordinates (with Cartesian axes in this book) in which positions and velocities of particles are measured.

  • Kinetic energy (½mv²) depends on velocity, so it depends on the choice of reference frame.
  • The excerpt emphasizes that energy measurements require specifying which frame is used.

⚖️ Inertial and lab frames

Inertial frame: a reference frame in which Newton's laws of motion are obeyed.

Lab frame: a reference frame whose axes are fixed relative to the earth's surface; for practical purposes it is inertial.

  • The laws of thermodynamics have been experimentally validated in stationary (lab) frames.
  • An inertial frame could be fixed or moving at constant velocity relative to local stars.
  • Example: a system in outer space can use an inertial frame fixed or moving uniformly; a system on earth typically uses an earth-fixed lab frame.

🔋 System energy E_sys

  • E_sys is the sum of the energies of all particles in the system plus their interaction potential energies, measured in a specified inertial frame.
  • The conservation of energy principle: the sum of system energy, surroundings energy, and any shared energy (all in the same frame) remains constant over time.
  • If the system as a whole moves or rotates relative to the inertial frame, E_sys depends on coordinates that are not properties of the system alone.
  • In such cases, E_sys is not a state function.

🧩 Internal energy U

🧩 Definition and purpose

Internal energy U: the energy of the system measured in a reference frame (called a local frame) that allows U to be a state function—at each instant, U depends only on the state of the system.

  • The local frame is chosen so that U does not include contributions from the system's motion or position as a whole.
  • U is always a state function; E_sys may or may not be, depending on whether the system moves in the lab frame.

🧪 Water-in-beaker example

The excerpt uses water in a glass beaker (the glass is part of the surroundings) to illustrate the distinction:

ScenarioWhat happens to E_sysWhat happens to UWhy
Beaker at rest on benchE_sys = UU depends on T and pLocal frame = lab frame
Heat water on hot plateE_sys and U both increaseU increasesTemperature T rises, changing the state
Slide beaker horizontally (T, p constant)E_sys increasesU unchangedKinetic energy in lab frame increases, but state (T, p) is the same
Lift beaker slowly (T, p constant)E_sys increasesU unchangedPotential energy in earth's gravity increases, but state is unchanged
  • The relation shown: ΔE_sys = ΔE_k + ΔE_p + ΔU, where E_k and E_p are kinetic and potential energies of the system as a whole in the lab frame.
  • Don't confuse: moving or lifting the system changes E_sys but not U if the thermodynamic state (T, p, etc.) is unchanged.

🎚️ Choosing a local frame

The excerpt lists three possible choices for the local frame:

  1. Lab frame: If the system does not move or rotate in the lab, a lab frame is appropriate. Then U = E_sys.
  2. Center-of-mass frame: If the system's center of mass moves in the lab, use a frame whose origin moves with the center of mass and whose axes are parallel to the lab axes.
  3. Container-fixed frame: If the system is in a rigid container that moves or rotates, use a frame fixed to the container.
  • The choice is "to some extent arbitrary" but should ensure U is a state function.

📏 Measuring internal energy

📏 Why absolute values are impractical

  • In principle, the total energy of a body at rest is E = mc² (Einstein's relation, where c is the speed of light).
  • So one could calculate U from the system's mass.
  • In practice, this is useless: the typical uncertainty in mass measurement (about 0.1 μg from a microbalance) translates to an enormous energy uncertainty of about 10¹⁰ joules.
  • Mass changes during ordinary chemical processes are far too small to detect.

📏 Only changes matter

  • Only values of the change ΔU are useful in thermodynamics.
  • ΔU cannot be calculated from the mass change Δm because Δm is undetectably small.
  • The excerpt emphasizes that absolute values of U have "too much uncertainty to be of any practical use."

🔗 Connection to the first law

🔗 Preview of Chapter 3

The excerpt briefly introduces the first law of thermodynamics (covered in Chapter 3):

  • Differential form: dU = ∂q + ∂w
  • Integrated form: ΔU = q + w

where:

  • U is the internal energy (a state function)
  • q is heat
  • w is thermodynamic work

🔗 Heat and work as energy transfer modes

Heat: transfer of energy across the boundary caused by a temperature gradient at the boundary.

Work: (definition cut off in the excerpt, but it is another mode of energy transfer).

  • Heat and work are path functions; their values depend on the process, not just initial and final states.
  • The first law states that the change in internal energy equals the sum of heat and work for a closed system.
10

Heat, Work, and the First Law

3.1 Heat, Work, and the First Law

🧭 Overview

🧠 One-sentence thesis

The first law of thermodynamics establishes that the change in a system's internal energy equals the sum of heat and work transferred across its boundary, where heat and work are path-dependent energy transfers rather than properties of the system itself.

📌 Key points (3–5)

  • The first law equation: ΔU = q + w, where U is internal energy (a state function), q is heat, and w is work (both path functions).
  • Heat vs work definitions: heat is energy transfer caused by a temperature gradient at the boundary; work is energy transfer caused by displacement under force or other directed motion.
  • Sign conventions: positive q or w increases internal energy (energy enters the system); negative q or w decreases it (energy leaves the system).
  • Common confusion: heat and work are not contained in a system—they are path-dependent transfers that occur during a process, not state functions like internal energy.
  • Path dependence: different processes with the same initial and final states can have different q and w values, but their sum ΔU is always the same.

🔋 The First Law Equations

🔋 Differential and integrated forms

The first law appears in two forms:

FormEquationMeaning
DifferentialdU = δq + δwInfinitesimal change in internal energy
IntegratedΔU = q + wFinite change in internal energy
  • δq and δw denote infinitesimal quantities of heat and work (not exact differentials).
  • To obtain finite q or w, integrate over the entire boundary surface and the entire process path.
  • The equation applies only to closed systems (no matter crosses the boundary).

⚖️ Energy conservation principle

The principle of conservation of energy: the total energy (sum of system and surroundings energies) remains constant over time.

  • When energy enters the system as heat, the same quantity leaves the surroundings.
  • When surroundings perform work on the system, the energy increase in the system equals the energy decrease in the surroundings.
  • Energy transfers are quantitative—amounts are conserved across the boundary.

🔥 Heat Transfer

🔥 What heat means

Heat: energy transfer across the boundary caused by a temperature gradient at the boundary.

  • Heat transfer occurs by conduction, convection, or radiation.
  • Positive heat (q > 0): energy entering the system.
  • Negative heat (q < 0): energy leaving the system.
  • Don't confuse: "heat" (energy transfer) with "heating" (causing temperature to increase)—heating can also be done by work alone.

🧊 Adiabatic processes

  • To completely prevent heat during a process, arrange conditions so there is no temperature gradient at any part of the boundary.
  • Under these conditions the process is adiabatic, and any energy transfer in a closed system is solely by work.
  • Practical methods to reduce heat: thermal insulation (reduces conduction), vacuum gap (eliminates conduction and convection), reflective surfaces (minimizes radiation).

🌡️ Heat capacity

Heat capacity: the ratio of an infinitesimal quantity of heat transferred under specified conditions to the resulting infinitesimal temperature change.

  • Definition: heat capacity = δq / dT (for a closed system).
  • Because q is a path function, heat capacity depends on specified conditions.
  • C_V: heat capacity at constant volume.
  • C_p: heat capacity at constant pressure.
  • Both are extensive state functions.

🔴 Common confusion: heat vs heating vs thermal energy

  • Heat (noun): energy transferred across the boundary due to a temperature gradient.
  • Heating (verb): causing a system's temperature to increase (can be done by heat or work).
  • Thermal energy: kinetic energy of random molecular motions plus vibrational and rotational energies; a contribution to internal energy.
  • Don't confuse: a change in thermal energy is not the same as heat unless the system is closed, there is no work, no volume change, no phase change, and no chemical reaction.
  • Example: everyday phrases like "heat flows from hot to cold" can mislead—heat is not a substance that retains identity; it is an energy transfer mode.

⚙️ Work Transfer

⚙️ What work means

Work: energy transfer across the boundary caused by displacement of a macroscopic portion of the system under force, or by other kinds of concerted, directed movement of entities (e.g., electrons) under external force.

  • Positive work (w > 0): work done by surroundings on the system (energy enters).
  • Negative work (w < 0): work done by system on surroundings (energy leaves).
  • Work involves short-range contact forces at the boundary.
  • Don't include: forces from conservative time-independent external fields (e.g., gravity)—these cause equal and opposite changes in potential and kinetic energy with no net effect on internal energy.

🧮 Evaluating work

For linear displacement in the x direction:

  • δw = F_sur,x dx or w = ∫ F_sur,x dx (from x₁ to x₂)
  • F_sur,x is the force component exerted by surroundings on the system.
  • By Newton's third law, can also write: δw = −F_sys,x dx, where F_sys,x is the force exerted by the system on the surroundings.

Alternative method: imagine the only effect on surroundings is a change in elevation of a weight linked mechanically to the force source. Then w = mgΔh (mass × gravity × height change). This helps determine sign and whether work occurs, but cannot determine the value unless an actual weight is present.

🔧 Work coefficients and work coordinates

For a single kind of work:

  • δw = Y dX or w = ∫ Y dX (from X₁ to X₂)
  • Y: generalized force (work coefficient).
  • X: generalized displacement (work coordinate).
  • Y and X are conjugate variables.
  • Example (from later in the text): reversible expansion work is δw = −p dV, where work coefficient is −p and work coordinate is V.

For multiple kinds of work:

  • δw = Σᵢ Yᵢ dXᵢ or w = Σᵢ ∫ Yᵢ dXᵢ

🌍 Local frame vs lab frame

  • The first law ΔU = q + w is measured in an arbitrary local frame.
  • Analogous relation in a stationary lab frame: ΔE_sys = q_lab + w_lab.
  • If the local frame's axes do not rotate relative to the lab frame, then q = q_lab (heat is the same in both frames).
  • For a center-of-mass frame (origin moves with system's center of mass, no rotation): w = w_lab − ½m Δ(v_cm²) − mg Δz_cm, where v_cm is center-of-mass velocity and z_cm is height in the lab frame.
  • In typical processes, v_cm and z_cm change negligibly, so w ≈ w_lab.
  • Exception: local frames with rotational motion (e.g., centrifuge cell)—simple relations do not exist; discussed later in the text.

🛤️ Path Dependence of Heat and Work

🛤️ The three-experiment demonstration

The excerpt describes three experiments with a system (water + paddle wheel + resistor) that start at T₁ = 300.0 K and end at T₂ = 300.10 K:

ExperimentProcessEnergy transferResult
1Release weight → paddle wheel churns water (insulated)Mechanical work onlyΔT = 0.10 K
2Close switch → electrical current through resistor (insulated)Electrical work only (same magnitude as exp. 1)ΔT = 0.10 K
3Place in thermal contact with 300.10 K reservoir (no insulation)Heat only (no work)ΔT = 0.10 K
  • All three processes have the same initial and final states, but different paths.
  • The sum q + w is the same in all three (equals ΔU), but the individual values of q and w differ.
  • An observer who sees only initial and final states cannot tell which path was taken—cannot determine how much energy was transferred as heat vs work.

🚫 Heat and work are not state functions

  • A system does not "have" or "contain" a particular quantity of heat or work at a given instant.
  • Heat and work depend on the path of a process occurring over time—they are path functions.
  • Only the sum q + w (which equals ΔU) is path-independent.
  • Internal energy U is a state function—its change depends only on initial and final states, not on the path.

🔄 Spontaneous and Reversible Processes

🔄 Spontaneous processes

Spontaneous process: a process that can actually occur in a finite time period under the existing conditions.

  • Any change over time in a system's state that we observe experimentally is a spontaneous process.
  • Also called: natural process, feasible process, possible process, allowed process, or real process.

🔄 Reversible processes (introduction)

Reversible process: an important concept needed for deriving the existence of entropy as a state function and establishing criteria for spontaneity and equilibrium.

  • The excerpt introduces the concept but does not provide the full definition here.
  • States that reversible processes are needed for the reasoning in the next chapter (deriving entropy).
  • Reversible processes lead to useful equalities among heat, work, and state functions like Gibbs energy.
  • (Full treatment deferred to later sections of the text.)
11

Spontaneous, Reversible, and Irreversible Processes

3.2 Spontaneous, Reversible, and Irreversible Processes

🧭 Overview

🧠 One-sentence thesis

A reversible process is an idealized limit of infinitely slow processes that pass through equilibrium states, useful for deriving thermodynamic relations even though it never actually occurs in reality.

📌 Key points (3–5)

  • Spontaneous vs. reversible: spontaneous processes actually occur in finite time; reversible processes are imaginary limits approached when real processes become infinitely slow.
  • What makes a process reversible: the system passes through a continuous sequence of equilibrium states that can be approximated by sufficiently slow real processes.
  • Common confusion: reversible processes are not real processes—they are idealized limits that real processes approach as they slow down; equations for reversible processes describe slow real processes to high accuracy.
  • Why reversibility matters: the concept is needed to derive entropy as a state function and to establish criteria for spontaneity and equilibrium.
  • Heat and work in reverse: during a reversible process, if you reverse direction, the magnitudes of heat and work stay the same but their signs flip.

🔄 Spontaneous processes and their reverse

🔄 What spontaneous means

A spontaneous process is a process that can actually occur in a finite time period under the existing conditions.

  • Any change over time in a system's state that we observe experimentally is spontaneous.
  • Also called: natural process, feasible process, possible process, allowed process, or real process.
  • The key is that it actually happens in the real world within finite time.

↩️ The reverse of a process

If a process takes the system from initial state A through intermediate states to final state B, then the reverse is a change from B to A with the same intermediate states in reverse time sequence.

  • Visualization: imagine filming the process, then running the film backward.
  • Each frame shows a "snapshot" of the state at one instant.
  • When reversed: property values (like pressure p and volume V) change in reverse chronological order, and each velocity changes sign.

🚫 Irreversible processes

If a process is spontaneous, which implies its reverse cannot be observed experimentally, the process is irreversible.

  • Spontaneous processes are irreversible because their reverse does not actually occur.
  • Example: you can observe a process happening, but you cannot observe it spontaneously reversing itself.

🎯 The reversible process concept

🎯 Core definition

A reversible process is an idealized process with a sequence of equilibrium states that are those of a spontaneous process in the limit of infinite slowness.

  • Not a real process: it is a hypothetical limit that real processes approach.
  • During a reversible process, the system passes through a continuous sequence of equilibrium states.
  • These equilibrium states can be approached as closely as desired by making a real spontaneous process go sufficiently slowly.
  • The slower the process, the more time between successive states for equilibrium to be approached.

⏱️ The infinite slowness limit

  • As a spontaneous process is carried out more and more slowly, it approaches the reversible limit.
  • Fermi's description: "A transformation is said to be reversible when the successive states differ by infinitesimals from equilibrium states."
  • A reversible transformation can be realized in practice by changing external conditions so slowly that the system has time to adjust itself gradually.
  • Don't confuse: reversible does not mean "fast enough to reverse"—it means "slow enough to stay in equilibrium."

🧩 Equilibrium states

  • Each state in a reversible process is an equilibrium state: one that in an isolated system would persist with no tendency to change over time.
  • This kind of process is sometimes called a quasistatic process.
  • The system's temperature and pressure must be uniform throughout (though this is only an approximation in real processes).

🔧 Characteristics of reversible processes

✅ Six defining characteristics

A reversible process of a closed system has all of the following:

CharacteristicWhat it means
1. Imaginary equilibrium sequenceThe system passes through a continuous sequence of equilibrium states (quasistatic process).
2. Approximable by slow real processesThe equilibrium sequence can be approximated as closely as desired by a real spontaneous process carried out sufficiently slowly; the reverse sequence can also be approximated by another slow spontaneous process.
3. Controlled by surroundingsThe process must involve heat or work (not isolated); an experimenter can use surroundings to control the rate of energy transfer and make the process as slow as desired.
4. Finite work coefficientIf work is transferred, the work coefficient Y in the expression (work = Y times change in X) must be finite (nonzero) in equilibrium states.
5. No internal dissipationIn the reversible limit, any energy dissipation within the system (e.g., internal friction) vanishes.
6. Reversible heat and workWhen any infinitesimal step reverses, the magnitudes of heat and work are unchanged but their signs reverse.

🚫 What is excluded

  • No hysteresis: processes with hysteresis (like plastic deformation or stretching a metal wire beyond its elastic limit) cannot have a reversible limit.
  • No very slow spontaneous changes: during the approach to infinite slowness, very slow changes must be eliminated with hypothetical constraints.
  • No internal friction: energy dissipation within the system must vanish in the reversible limit.

🔁 Energy transfer in reverse

  • Energy transferred as heat in one direction during a reversible process is transferred as heat in the opposite direction during the reverse process.
  • The same is true for work.
  • Example: if 10 joules of heat enter the system during a reversible step, then 10 joules of heat leave the system when that step is reversed.

🎬 Imagining a reversible process

🎬 The mental picture

  • Imagine a gas whose volume, temperature, and pressure are changing at some finite rate.
  • The temperature and pressure "magically" stay perfectly uniform throughout the system.
  • This is entirely imaginary because there is no temperature or pressure gradient—no physical "driving force"—to make the change occur in a particular direction.
  • The states of uniform temperature and pressure are approached by a real process as the real process becomes slower and slower.

💡 Practical interpretation

  • Whenever you see "reversible," think "in the reversible limit."
  • Reversible process = process in the reversible limit.
  • Reversible work = work in the reversible limit.
  • Reversible heat = heat in the reversible limit.

📐 Why reversible processes matter

📐 Theoretical importance

  • The concept is needed for deriving the existence of entropy as a state function and defining its changes.
  • Entropy then leads to criteria for spontaneity and for various kinds of equilibria.
  • Innumerable useful relations (equalities) among heat, work, and state functions (like Gibbs energy) can be obtained for reversible processes.

📐 Practical use of reversible equations

  • Common confusion: "What is the use of an equation for a process that can never actually occur?"
  • Answer: The equation can describe a spontaneous process to a high degree of accuracy if the process is carried out slowly enough.
  • For many important spontaneous processes, we can assume temperature and pressure are uniform throughout the system, though this is only an approximation.
  • Example: if a real process is slow enough that intermediate states depart only slightly from exact equilibrium, a reversible equation describes it accurately.

🌍 Reversibility and the surroundings

🌍 The reverse of a reversible process

  • The reverse of a reversible process is itself a reversible process.
  • Energy transferred as heat and work during a reversible process is returned across the boundary when the process is followed by the reverse process.

🌍 Can surroundings be restored?

  • Some authors describe a reversible process as one that allows both system and surroundings to be restored to their initial states.
  • Problem with this description: during the time period of the process and its reverse, spontaneous irreversible changes inevitably occur in the surroundings.

🌍 Local vs. auxiliary surroundings

  • Zemansky and Dittman define:
    • Local surroundings: parts that interact directly with the system to transfer energy (e.g., a weight that does work, heat reservoirs in thermal contact).
    • Auxiliary surroundings (rest of the universe): parts that might interact with the system (e.g., mechanisms to move weights or heat reservoirs).
  • Issue: controlling external operations requires a human operator or automated mechanism whose actions are spontaneous and irreversible.
  • If these are part of the auxiliary surroundings, then not all auxiliary surroundings can return to their initial states.
  • Don't confuse: "reversible" does not mean "everything in the universe returns to its initial state"—it refers to the idealized limit of the system's process.

🌍 Example: cylinder and piston

  • The excerpt mentions a cylinder-and-piston device (Figure 3.2) to illustrate a reversible process whose reverse does not restore the local surroundings.
  • System: confined gas.
  • Local surroundings: piston (a weight) and heat reservoir.
  • (The excerpt does not provide further details of this example.)
12

Heat Transfer

3.3 Heat Transfer

🧭 Overview

🧠 One-sentence thesis

Heat transfer processes can be either irreversible (occurring at finite rates with temperature gradients) or reversible (infinitely slow with infinitesimal temperature differences), and the distinction determines whether the process can truly be reversed without producing impossible microscopic events.

📌 Key points (3–5)

  • What heat transfer means: energy crossing the boundary due to a temperature gradient, always flowing from higher to lower temperature.
  • Irreversible heat transfer: occurs at finite rates with finite temperature differences across the boundary, creating non-equilibrium states within the system.
  • Reversible heat transfer: requires infinitely slow processes with infinitesimal temperature differences, maintaining near-equilibrium states throughout.
  • Common confusion: a reversible process does not mean both system and surroundings can be restored without any changes elsewhere—the book defines reversibility by the system's internal characteristics alone (internal reversibility).
  • Why it matters: understanding reversible vs irreversible processes is fundamental to thermodynamics and explains why certain natural processes cannot spontaneously reverse.

🔄 Understanding reversible vs irreversible processes

🔄 What makes a process reversible

The excerpt defines a reversible process by characteristics of the system itself, not by whether surroundings can be restored:

  • The system passes through equilibrium states (or states differing only slightly from equilibrium)
  • The process occurs infinitely slowly
  • The process can be reversed by an infinitesimal change in conditions
  • This is called internal reversibility

A reversible process is defined by characteristics involving only changes in the system itself, regardless of what happens in the surroundings.

Don't confuse: Some authors claim a reversible process allows both system and surroundings to be restored, but the excerpt argues this is neither useful nor valid because spontaneous irreversible changes inevitably occur in the surroundings during any real time period.

🎯 The piston-cylinder example

The excerpt uses a gas confined by a piston to illustrate why "restoring the surroundings" is problematic:

Initial state: Gas at temperature T_res, volume V₁, pressure p₁; piston held at height h₁ by catches.

Expansion process:

  • Remove catches; piston rises slowly due to high-viscosity lubricant
  • Friction at the piston seal produces thermal energy transferred as heat to the reservoir
  • In the limit of infinite slowness, this approaches a reversible isothermal expansion
  • Gas reaches new equilibrium: same temperature T_res, larger volume, lower pressure

Compression (reverse) process:

  • Add weight to piston; it sinks slowly back to height h₁
  • Again friction produces heat transferred to the reservoir
  • In the limit of infinite slowness, this is a reversible isothermal compression
  • System returns to initial state (T = T_res, V = V₁, p = p₁)

Key observation: The system returned to its initial state, but the local surroundings did not—a weight was added to the piston, and the heat reservoir's internal energy increased due to friction. Restoring these would require further irreversible changes in auxiliary surroundings (mechanisms, operators).

⚠️ What makes a process irreversible

An irreversible process is a spontaneous process whose reverse is neither spontaneous nor reversible—the reverse can never actually occur and is impossible.

The movie test: If you film a spontaneous process and run the film backward, and the reversed sequence could not occur in reality, the process is irreversible.

Example from the excerpt: A sinking weight drives a paddle wheel in water, increasing the water's temperature (mechanical energy dissipates into thermal energy). Running this film backward shows the paddle wheel raising the weight while the water cools—impossible in reality because H₂O molecules would need to move in concerted motion simultaneously, which is extremely unlikely.

🎾 Purely mechanical processes (special case)

A third category exists: processes that are spontaneous in both directions.

Purely mechanical processes involve motion of perfectly-elastic macroscopic bodies without friction, temperature gradients, viscous flow, or other irreversible changes.

Examples:

  • A ball thrown through a vacuum (can move spontaneously left or right)
  • A frictionless flywheel rotating in a vacuum

Important notes:

  • These are idealizations of a different kind than reversible processes
  • They proceed at finite rates, so their states are not equilibrium states
  • They are not reversible processes
  • They are of little interest in chemistry
  • Later chapters treat "spontaneous" and "irreversible" as synonyms, ignoring this special case

🌡️ Heat transfer mechanisms

🌡️ Basic definition and direction

Heat transfer or heat flow: energy being transferred across the boundary on account of a temperature gradient at the boundary.

Key characteristics:

  • Transfer is always in the direction of decreasing temperature
  • From warmer side to cooler side across the boundary
  • The temperature may appear discontinuous at the boundary (different values on either side), but actually there is a thin zone with a temperature gradient

🔥 Irreversible heating example: metal sphere in water bath

The excerpt illustrates irreversible heat transfer with a solid metal sphere immersed in a well-stirred, temperature-controlled water bath.

Setup:

  • Initial equilibrium: sphere and bath both at T₁ = 300.0 K
  • Goal: raise sphere temperature by one kelvin to T₂ = 301.0 K

Irreversible method:

  • Rapidly increase bath temperature to 301.0 K and hold it there
  • Temperature difference across the sphere's surface causes spontaneous heat flow into the sphere
  • Heat flows through the system boundary into the sphere

Why it's irreversible:

  • Takes time for all parts of the sphere to reach the higher temperature
  • Creates temporary non-equilibrium states (temperature gradients within the sphere)
  • The process occurs at a finite rate with a finite temperature difference (1 K)

📊 Temperature profiles during heating

The excerpt references Figure 3.4, which shows temperature profiles in a copper sphere (radius 5 cm) at different times during heating:

TimeDescription
Initial stateUniform temperature throughout sphere
2 s, 4 s, 6 s, 8 s, 10 sTemperature varies with distance r from center (non-equilibrium)
Final stateUniform temperature again (new equilibrium)

Key insight: During the process, temperature is a function of position (distance r from center), showing the system passes through non-equilibrium states—a hallmark of irreversible heat transfer.

🔬 Approaching reversible heat transfer

🐌 The requirement for reversibility

To make heat transfer reversible, the process must be infinitely slow with infinitesimal temperature differences:

  • The temperature difference between bath and sphere must approach zero
  • The process must proceed so slowly that the sphere remains arbitrarily close to equilibrium at all times
  • In the limit, every intermediate state is an equilibrium state
  • This is an idealization that cannot be achieved in practice but serves as a theoretical benchmark

Don't confuse: "Reversible" does not mean "fast enough to reverse easily"—it means the opposite: infinitely slow so that the system is always in equilibrium and an infinitesimal change in conditions can reverse the direction.

🔄 Contrast: reversible vs irreversible heat transfer

AspectIrreversibleReversible
Temperature differenceFinite (e.g., 1 K)Infinitesimal (approaches zero)
RateFinite (occurs in real time)Infinitely slow
System statesNon-equilibrium (gradients within)Equilibrium (or infinitesimally close)
PracticalAchievable in realityTheoretical idealization
ExampleSphere in bath with ΔT = 1 KSphere in bath with ΔT → 0
13

Deformation Work

3.4 Deformation Work

🧭 Overview

🧠 One-sentence thesis

Deformation work in a gas can be carried out either irreversibly (with pressure gradients and friction during spontaneous expansion or compression) or reversibly (in the limit of infinite slowness, passing through equilibrium states with uniform pressure and temperature).

📌 Key points (3–5)

  • What deformation work is: work done when the system boundary changes position, such as when a gas expands or compresses in a cylinder-and-piston device.
  • Spontaneous processes are irreversible: rapid expansion or compression creates pressure and temperature gradients inside the gas, so intermediate states are not equilibrium states.
  • Reversible limit: when the process occurs infinitely slowly, the gas passes through a continuous sequence of equilibrium states with uniform temperature and pressure.
  • Common confusion: expansion vs compression gradients—the pressure gradient during spontaneous expansion is opposite in direction to the gradient during spontaneous compression, so neither can occur in reverse chronological order.
  • Pressure at the moving boundary: the pressure at the piston differs slightly from the pressure at the stationary wall during motion, but this difference becomes negligible at low piston speeds.

🔧 What is deformation work

🔧 General definition

Deformation of a system involves changes in the position, relative to the local frame, of portions of the system boundary.

  • At a small surface element of the boundary, the work of deformation is given by: the magnitude of the contact force exerted by the surroundings on the surface element, times the cosine of the angle between force and displacement, times the infinitesimal displacement.
  • If the displacement is entirely parallel to one axis (x), the expression simplifies to: force in the x direction times displacement in x.

🎯 Volume change as a key example

  • A useful kind of deformation for thermodynamic theory is a change in the volume of a gas or liquid.
  • The excerpt uses a gas confined in a horizontal cylinder by a piston as the model system.
  • The system is the gas; the piston is not part of the system, but its position determines the system's volume.

🔩 Forces and pressure in the cylinder-and-piston device

🔩 Three kinds of forces on the piston

The excerpt identifies three forces acting on the piston:

ForceDirectionDescription
F_gas+x (rightward)Force exerted by the gas on the piston; equals pressure at the piston times the cross-section area
F_ext−x (leftward)External force controlled in the surroundings
F_fricOpposite to velocityFrictional force at the seal; approaches zero as piston velocity approaches zero
  • The net force on the piston is: F_gas minus F_ext plus F_fric.
  • When the piston is stationary and the gas is in equilibrium at uniform pressure p₁, the net force is zero, so F_ext equals F_gas equals p₁ times area.

📐 Pressure at the boundary vs the wall

  • p_b: the average pressure of the gas at the piston (the moving portion of the system boundary).
  • p: the pressure at the stationary cylinder wall.
  • During motion, p_b differs slightly from p because of the piston's velocity relative to the gas molecules.
  • The excerpt derives that p_b is related to p by a factor involving the ratio of piston speed to molecular velocities.
  • Example: for nitrogen at 300 K with piston speed 10 m/s, p_b is only about 5% lower than p during expansion and 5% higher during compression.
  • At low piston speeds, the percentage difference is proportional to speed and can usually be ignored for practical calculations.

🌀 Spontaneous expansion and compression (irreversible processes)

🌀 Spontaneous expansion

  • Start with the gas in an equilibrium state of uniform temperature T₁ and uniform pressure p₁, with the piston stationary.
  • To avoid heat transfer complications, the excerpt assumes an adiabatic boundary.
  • By reducing F_ext from its initial value, spontaneous expansion begins.
  • As the piston moves to the right, the pressure p_b at the piston becomes slightly less than the pressure at the stationary wall.
  • Molecular explanation: gas molecules moving to the right approach the moving piston at lower velocities relative to the piston, so they collide less frequently and with smaller momentum loss per collision.
  • Temperature and pressure within the gas become nonuniform; intermediate states cannot be described with single values of T and p.
  • These intermediate states are not equilibrium states.

🌀 Spontaneous compression

  • By increasing F_ext from the value at the end of expansion, spontaneous compression begins.
  • The gas pressure p_b at the piston now becomes slightly greater than at the stationary wall, because the piston is moving toward the molecules that are moving to the right.
  • A different pressure gradient develops than during expansion.
  • Don't confuse: the pressure gradient during compression is in the opposite direction to the gradient during expansion.
  • It is physically impossible for the sequence of states of either process to occur in reverse chronological order, because that would have thermal energy (or momentum transfer) flowing in the wrong direction along the gradient.
  • This shows that spontaneous expansion and compression are irreversible.

⚖️ The reversible limit

⚖️ Infinitely slow processes

  • The more slowly we allow the adiabatic expansion (or compression) to take place, the more nearly uniform are the temperature and pressure.
  • In the limit of infinite slowness, the gas passes through a continuous sequence of equilibrium states of uniform temperature and pressure.
  • The states approached in the limit of infinitely slow compression are equilibrium states, occurring in the reverse sequence of the states for infinitely slow expansion.

⚖️ Definition of a reversible process

The sequence of equilibrium states, taken in either direction, is a reversible process.

  • In the reversible limit, the heating and cooling processes (or expansion and compression processes) have the same intermediate states.
  • These states have no temperature gradients (or pressure gradients).
  • Example: let p₂ be the pressure in the final state of the infinitely-slow expansion; in this state, F_ext equals p₂ times area. By increasing F_ext from this value, we can reverse the process and cause compression through the same equilibrium states.

⚖️ Why reversibility matters

  • Only in the reversible limits do the expansion and compression processes share the same intermediate states.
  • Spontaneous (irreversible) processes have different gradients in opposite directions, so they cannot be reversed by simply running time backward.
  • The excerpt emphasizes that the intermediate states of spontaneous processes are not equilibrium states, whereas the reversible limit consists entirely of equilibrium states.
14

Applications of Expansion Work

3.5 Applications of Expansion Work

🧭 Overview

🧠 One-sentence thesis

By controlling the speed of expansion or compression, a gas can be made to pass through either a sequence of equilibrium states (reversible process) or non-equilibrium states (spontaneous process), with the pressure at the moving piston differing only slightly from the pressure at the stationary wall even at moderate piston speeds.

📌 Key points (3–5)

  • Reversible vs spontaneous processes: infinitely slow expansion/compression allows the gas to pass through equilibrium states; faster processes create non-uniform temperature and pressure (non-equilibrium states).
  • Pressure gradient during motion: when the piston moves, the pressure at the piston face differs slightly from the pressure at the stationary wall due to molecular collision dynamics.
  • Direction matters: during expansion, pressure at the piston is slightly less than at the wall; during compression, it is slightly greater.
  • Common confusion: the pressure difference is small—even at 10 m/s piston speed, the difference is only about 5% for nitrogen at 300 K, and decreases proportionally with slower speeds.
  • Equilibrium limit: the slower the process, the more uniform the temperature and pressure become; infinite slowness yields a reversible process.

🔄 Reversible vs spontaneous processes

🛑 Initial equilibrium state

  • The gas starts with uniform temperature T₁ and uniform pressure p₁.
  • The piston is stationary, so friction force F_fric is zero.
  • Net force F_net is zero (Newton's second law), so the external force F_ext equals the gas force: F_ext = p₁A_s (where A_s is piston area).
  • The system has an adiabatic boundary (no heat transfer).

⚡ Spontaneous expansion

  • Reducing F_ext below p₁A_s causes the piston to move to the right (expansion begins).
  • As the piston moves, temperature and pressure become non-uniform within the gas.
  • Intermediate states cannot be described by single values of T and p—they are not equilibrium states.

Spontaneous process: a process in which the system passes through non-equilibrium states with non-uniform properties.

🐢 Infinitely slow expansion (reversible process)

  • The slower the expansion, the more nearly uniform the temperature and pressure remain.
  • In the limit of infinite slowness, the gas passes through a continuous sequence of equilibrium states with uniform T and p.
  • Final pressure p₂ corresponds to F_ext = p₂A_s.

Reversible process: a sequence of equilibrium states that can be traversed in either direction (expansion or compression).

  • Don't confuse: reversible does not mean "undoable"; it means the process passes through equilibrium states and can be run forward or backward through the same sequence.

🔁 Spontaneous compression

  • Increasing F_ext above p₂A_s causes the piston to move left (compression begins).
  • Pressure at the piston p_b becomes slightly greater than at the stationary wall (opposite gradient from expansion).
  • In the limit of infinite slowness, compression passes through the same equilibrium states as expansion, but in reverse order.

🎯 Pressure gradient at the moving piston

🧪 Molecular explanation during expansion

  • Gas molecules moving to the right approach the moving piston at lower relative velocities than if the piston were stationary.
  • Result: molecules collide with the piston less frequently and with smaller momentum loss per collision.
  • Therefore, pressure p_b at the piston face is slightly less than pressure p at the stationary cylinder wall.

🧪 Molecular explanation during compression

  • The piston moves left toward molecules moving right.
  • Molecules collide with the piston at higher relative velocities.
  • Therefore, pressure p_b at the piston face is slightly greater than pressure p at the stationary wall.
ProcessPiston motionRelative collision velocityPressure at piston vs wall
ExpansionRight (away from gas)Lowerp_b < p
CompressionLeft (into gas)Higherp_b > p

🔬 Kinetic-molecular theory estimate

🧮 Collision dynamics

  • Consider a molecule of mass m colliding with the piston moving at constant velocity u (positive for expansion, negative for compression).
  • In the piston-fixed reference frame, the collision reverses the sign but not the magnitude of the velocity component: v'_x,2 = –v'_x,1.
  • In the lab frame, the velocity after collision is: v_x,2 = v_x,1 + 2u.

📐 Force and pressure calculation

  • The time-average force exerted by one molecule on the piston over one collision cycle is:
    • h F_x i = (m/l) × (v²_x,1 – 2u v_x,1)
    • where l is the interior cylinder length.
  • Summing over all nM/m molecules (n = amount, M = molar mass) and dividing by piston area A_s gives the pressure p_b at the piston.
  • Pressure at the stationary wall p is found by setting u = 0.

📊 Quantitative result

The relationship between p_b and p is:

p_b = p × [1 – 2u⟨v_x,1⟩ / ⟨v²_x,1⟩]

  • From kinetic-molecular theory: ⟨v_x,1⟩ = (2RT/M)^(1/2) and ⟨v²_x,1⟩ = RT/M.
  • Example: nitrogen (N₂) at 300 K, piston speed 10 m/s:
    • During expansion: p_b is about 5% lower than p.
    • During compression: p_b is about 5% higher than p.
  • At low piston speeds, the percentage difference is proportional to piston speed.

✅ Practical implication

  • Even at the considerable speed of 10 m/s, the pressure difference is small (only ~5%).
  • For reasonably slow speeds, the difference is quite small and can usually be ignored in practical calculations.
  • Don't confuse: this does not mean there is no difference—it means the difference is negligible for most purposes when the piston moves slowly.

🔧 Expansion work context

🏗️ Cylinder-and-piston device

  • The system is the gas inside the cylinder.
  • The piston has area A_s and can move horizontally.
  • The boundary is adiabatic (no heat transfer).
  • This setup is used to study expansion work (also called pressure-volume work), which applies to both expansion and compression.

🎓 Scope of analysis

  • The excerpt focuses on the pressure gradient and the conditions for reversibility.
  • The kinetic-molecular theory calculation is noted as "not part of classical macroscopic thermodynamics" but is used to estimate the magnitude of the pressure difference.
  • The analysis confines attention to adiabatic boundaries to avoid complications from heat transfer.
15

Work in a Gravitational Field

3.6 Work in a Gravitational Field

🧭 Overview

🧠 One-sentence thesis

The work done in a gravitational field depends critically on how the system boundary is defined, and reversible processes (controlled by external forces like a string) achieve the algebraically smallest work for a given displacement compared to spontaneous processes like free fall.

📌 Key points (3–5)

  • System boundary matters: defining the system as "body only" versus "body plus fluid" leads to different work calculations, differing by buoyancy and friction terms.
  • Reversible limit: when a body is lowered on a string at near-zero velocity, friction vanishes and work approaches the reversible minimum w = (mg − F_buoy)Δz.
  • Free fall is spontaneous: a freely falling body cannot be controlled from the surroundings, so the process has no reversible limit.
  • Common confusion: gravitational force mg acts on the body but does not count as F_sur (the contact force from surroundings) in the work formula; only contact forces—buoyancy, friction, and string tension—contribute to F_sur.
  • General principle: for any adiabatic process with a given change in the work coordinate, work is algebraically smallest (least positive or most negative) in the reversible limit, meaning the system does maximum work on the surroundings or the surroundings do minimum work on the system.

🎯 System definition and work calculation

🎯 Why system boundaries change the work

  • The excerpt emphasizes two valid choices for defining the system when a spherical body moves through a fluid in a gravitational field.
  • Choice affects which forces count as "surroundings acting on the system" and thus which terms appear in the work integral.

🧱 System = body + fluid

  • The system boundary is where the fluid contacts the vessel walls and atmosphere.
  • Because the vessel is stationary, there is no displacement of the boundary → no work (¶w = 0).
  • Example: a marble falling through water in a sealed container—if you treat marble + water as the system, the boundary (container walls) doesn't move, so w = 0.
  • If the process is adiabatic, internal energy stays constant: gravitational potential energy lost by the body converts to kinetic and thermal energy within the system.

🧱 System = body only

  • The fluid is now in the surroundings.
  • Contact forces on the body include:
    • Buoyant force F_buoy = ρV₀g (upward, where ρ is fluid density and V₀ is body volume).
    • Frictional drag F_fric (opposes velocity).
    • String tension F_str (if present).
  • Key point: the gravitational force −mg is not included in F_sur; only contact forces from the surroundings count.
  • Work formula: ¶w = F_sur_z dz, where F_sur_z is the net upward contact force.

🪂 Free fall through a fluid

🪂 Forces and work in free fall

  • When the body falls freely (no string), the only contact forces are buoyancy and friction.
  • Work done on the body (system = body only):

    ¶w = (−F_buoy + F_fric) dz

  • Because dz is negative (body falls), the work ¶w is negative: the body does work on the fluid.
  • The magnitude |F_buoy dz| is the work of pushing displaced fluid upward; |F_fric dz| is energy dissipated as heat by friction.

🪂 No reversible limit for free fall

  • The rate of energy transfer cannot be controlled from the surroundings.
  • Velocity and friction cannot be made to approach zero while still having the process occur.
  • Example: dropping a marble into water—you cannot slow it down from outside once released, so it is inherently irreversible.

🧵 Lowering on a string (controlled process)

🧵 String provides external control

  • A thin string attached to the body allows the surroundings to exert an upward contact force F_str.
  • The string is in the surroundings; the moving boundary is at the attachment point.
  • Newton's second law for the body:

    (−mg + F_buoy + F_fric + F_str) = m dv/dt

  • Solving for F_str:

    F_str = −mg − F_buoy − F_fric + m dv/dt

🧵 Work when system = body + fluid

  • Moving boundary is where the string attaches to the body.
  • Work: ¶w = F_str dz.
  • Substituting the expression for F_str:

    ¶w = (−mg − F_buoy − F_fric + m dv/dt) dz

  • The term (m dv/dt) dz is an infinitesimal change in kinetic energy dE_k.
  • For a process starting and ending in equilibrium (ΔE_k = 0), the finite work is:

    w = (−mg − F_buoy)Δz − ∫ F_fric dz

🧵 Reversible limit

  • As velocity v approaches zero, friction F_fric approaches zero.
  • Reversible work:

    w = (mg − F_buoy)Δz

  • If the fluid is a gas (density much smaller than body density), F_buoy ≈ 0, so:

    w ≈ mgΔz

  • Key insight: because F_fric and dz have opposite signs, the integral ∫ F_fric dz is always positive when the body is lowered (dz < 0, F_fric > 0) or negative when raised (dz > 0, F_fric < 0). Thus w is algebraically smallest (least positive or most negative) in the reversible limit.

🧵 Work when system = body only

  • F_sur_z = (−F_buoy + F_fric + F_str).
  • Substituting from Newton's second law:

    F_sur_z = (mg + m dv/dt)

  • Work:

    ¶w = (mg + m dv/dt) dz

  • For a process with ΔE_k = 0:

    w = mgΔz

  • Important: this work is independent of velocity during the process.
  • Reversible work (same as finite work when ΔE_k = 0):

    ¶w = mg dz and w = mgΔz

🧵 Comparing the two system definitions

  • When system = body only, work is greater by the amount (−F_buoy + F_fric) dz than when system = body + fluid.
  • This difference accounts for the work of displacing fluid and energy dissipated by friction.
  • Don't confuse: the "extra" work when system = body only is not "lost"—it reflects energy transferred to the fluid (buoyancy work) and converted to heat (friction).

🔄 General principle for adiabatic processes

🔄 Minimum work in the reversible limit

  • The excerpt states a general rule: for an adiabatic expansion or compression with a given change of the work coordinate, starting from a given initial equilibrium state, work is algebraically smallest in the reversible limit.
  • "Algebraically smallest" means:
    • Least positive (if work is done on the system).
    • Most negative (if work is done by the system).
  • Equivalently:
    • Surroundings do the least possible work on the system.
    • System does the maximum possible work on the surroundings.

🔄 Why this matters

  • Reversible processes are the benchmark for efficiency.
  • Any spontaneous (irreversible) process with the same initial state and same change in work coordinate will have algebraically larger work (more positive or less negative).
  • Example: lowering a body slowly on a string (reversible) requires less work input from surroundings than letting it fall and then lifting it back (spontaneous + reversal).

📐 Summary of work formulas

System definitionProcessWork formula (finite, ΔE_k = 0)Notes
Body + fluidFree fallw = 0Boundary doesn't move
Body onlyFree fallw = −∫(F_buoy − F_fric) dzNegative (body does work on fluid)
Body + fluidLowered on stringw = (−mg − F_buoy)Δz − ∫ F_fric dzReversible limit: w = (mg − F_buoy)Δz
Body onlyLowered on stringw = mgΔzIndependent of velocity

📐 Reversible gravitational work of a body

¶w = mg dz and w = mgΔz (reversible gravitational work of a body)

  • This formula applies when the system is the body only, in the reversible limit.
  • It is the simplest and most commonly used expression for gravitational work.
16

Shaft Work

3.7 Shaft Work

🧭 Overview

🧠 One-sentence thesis

Shaft work—energy transferred by a rotating shaft—can have a reversible limit when friction is minimized (as in lifting a weight), but stirring work is inherently dissipative because it converts mechanical energy into thermal energy regardless of rotation direction.

📌 Key points (3–5)

  • What shaft work is: energy transferred across a system boundary by a rotating shaft, calculated from the internal torque and angle of rotation.
  • Two kinds of shaft work: lifting work (System A, with a weight and pulley) has a reversible limit as angular velocity approaches zero; stirring work (System B, with a paddle in water) is always dissipative.
  • Dissipative work: work that is positive for both positive and negative changes of the work coordinate, so it cannot be reversed—energy is converted to thermal energy (heat) inside the system.
  • Common confusion: the work coordinate for stirring (rotation angle θ) is not a state function, unlike work coordinates for reversible work (e.g., volume V or height z).
  • Historical significance: Joule's paddle-wheel experiments quantified the "mechanical equivalent of heat" by measuring how much stirring work produces the same temperature rise as a given amount of heat.

🔧 Definition and general formula

🔧 What shaft work measures

Shaft work: energy transferred across the boundary by a rotating shaft.

  • The shaft passes through the system boundary and rotates through an angle θ (in radians).
  • Angular velocity ω = dθ/dt.
  • Torques act at both ends: τ_sys (inside the system) and τ_sur (in the surroundings).

📐 The shaft work formula

When angular velocity ω is zero or constant (no angular acceleration), the torques balance: τ_sys = −τ_sur.

  • An internal torque τ_b acts at the boundary cross-section of the shaft, with magnitude equal to τ_sys.

  • The shaft work is given by:

    w = integral from θ₁ to θ₂ of τ_b dθ = − integral from θ₁ to θ₂ of τ_sys dθ
    (shaft work, constant ω)

  • This formula applies when the shaft rotates at constant angular velocity (or is stationary).

🏋️ System A: Lifting work with a reversible limit

🏋️ Setup and torque

  • System A (Figure 3.11): a weight of mass m hangs from a cord wound around a pulley wheel attached to the shaft.
  • When ω = 0 (shaft at rest), the torque τ_sys = −mgr, where r is the radius at the cord attachment point.
  • When ω is finite and constant, friction at the shaft and pulley bearings modifies τ_sys:
    • If ω > 0 (shaft turning in the positive direction), friction makes τ_sys more negative than −mgr.
    • If ω < 0 (shaft turning in the negative direction), friction makes τ_sys less negative than −mgr.

🔄 Reversible limit

  • As ω → 0 (infinite slowness), frictional forces vanish.

  • The shaft work approaches the reversible limit:

    w = mgr Δθ
    (reversible shaft work, System A)

  • Key property: shaft work is least positive or most negative in the reversible limit.

  • Figure 3.12(a) shows how w varies with ω for a fixed rotation magnitude |θ₂ − θ₁|: the curve has a minimum (most negative work) at ω = 0.

🧩 Why this work is reversible

  • The weight can be raised or lowered by controlling the shaft velocity.
  • In the limit of zero velocity, the process can be reversed with the same magnitude of work (opposite sign).
  • The work coordinate (rotation angle θ) and the work coefficient (torque) are both related to state functions (the height of the weight).

🌀 System B: Stirring work (dissipative)

🌀 Setup and torque

  • System B (Figure 3.11): a stirrer (paddle) is immersed in water.
  • When ω = 0 and the water is at rest, τ_sys = 0 and τ_b = 0.
  • When ω is finite and constant, the stirrer creates turbulent motion in the water.
    • Frictional drag at the stirrer blades and shaft bearings produces a torque τ_sys with the opposite sign from ω.
    • The magnitude of τ_sys increases with the magnitude of ω.

🔥 Dissipative nature

  • For a fixed rotation magnitude |θ₂ − θ₁|, the stirring work depends on ω as shown in Figure 3.12(b):
    • Work is positive for both positive and negative ω.
    • Work approaches zero as ω → 0 (infinite slowness).
  • This is the defining feature of dissipative work:

    Dissipative work is work that is positive for both positive and negative changes of the work coordinate, and therefore cannot be carried out reversibly.

⚡ Energy conversion

  • Energy transferred by stirring work is converted by friction into random thermal motion (thermal energy) within the system.
  • The energy is completely dissipated—it cannot be recovered as work when the rotation is reversed.
  • Equivalence to heat: because stirring work is converted to thermal energy, it can be replaced by an equal quantity of positive heat to produce the same overall change.
    • Example: Experiment 3 (page 62) demonstrated this equivalence.

🚫 Not a state function

  • The rotation angle θ is a property of the system but not a state function.
    • The system can be in exactly the same state at θ = 0 and θ = 2π (one full rotation).
  • General rule: the work coordinate of any dissipative work is not a state function, whereas work coordinates for reversible work (e.g., volume V, height z) are state functions.

🔀 Multiple work coordinates

  • System B can have both stirring work (coordinate θ) and expansion work (coordinate V, with đw = −p dV).
  • Only the expansion work can be carried out reversibly.
  • The number of independent variables for equilibrium states (e.g., T and V, so 2 variables) is one greater than the number of work coordinates for reversible work (only V, so 1 coordinate).
    • This agrees with the general rule on page 77.

🔬 Joule's paddle-wheel experiments

🔬 Purpose and setup

  • Goal: determine the "mechanical equivalent of heat"—how much mechanical (stirring) work produces the same temperature increase as a given amount of heat.
  • Apparatus (Figures 3.13 and 3.14):
    • Eight sets of metal paddle arms attached to a rotating shaft inside a water-filled copper vessel.
    • Four sets of stationary metal vanes fixed inside the vessel, with openings for the paddle arms to pass through.
    • The vanes prevent the water from simply rotating in a circle, forcing turbulent (viscous) flow.
    • Two lead weights sink and cause the shaft to rotate via connecting strings and pulleys.

📏 Measuring stirring work

  • The system is the vessel, its contents, and the lid.

  • First approximation: stirring work equals the negative of the change in the weights' potential energy:

    w = mg Δh
    (where m is the combined mass of the two weights, Δh is the change in vertical position)

  • Corrections Joule made:

    • Kinetic energy gained by the weights.
    • Friction in the connecting strings and pulley bearings.
    • Elasticity of the strings.
    • Heat gain from the surrounding air.

📊 Results

  • Joule performed 40 experiments at average temperatures of 13–16 °C.
  • Result: the work needed to increase the temperature of one gram of water by one kelvin was 4.165 J.
  • This value is close to the modern accepted value for the specific heat capacity of water.
  • Significance: these experiments provided quantitative evidence that mechanical work and heat are interconvertible forms of energy, driving the final rejection of the caloric theory.

🎯 Key insight

  • Stirring work is entirely converted to thermal energy (temperature increase) inside the system.
  • The same temperature change can be produced by adding heat instead of doing stirring work, demonstrating the equivalence of work and heat as forms of energy transfer.

🔍 Comparing reversible and dissipative shaft work

FeatureSystem A (lifting)System B (stirring)
Torque at ω = 0Finite (−mgr)Zero
Work vs ωMinimum at ω = 0 (reversible limit)Positive for both ±ω, zero at ω = 0
ReversibilityReversible in limit ω → 0Inherently irreversible (dissipative)
Energy fateCan be recovered as workConverted to thermal energy, dissipated
Work coordinateRelated to state function (height)Not a state function (rotation angle θ)
Sign of workCan be positive or negativeAlways positive for finite ω

🧠 Don't confuse

  • Reversible shaft work (System A): friction vanishes as ω → 0, and work can be reversed.
  • Dissipative shaft work (System B): even at very slow rotation, friction is essential to the process (it creates the turbulent flow), so work is always positive and cannot be reversed.
17

Electrical Work

3.8 Electrical Work

🧭 Overview

🧠 One-sentence thesis

Electrical work in thermodynamic systems depends on electric potential differences and charge flow, with different characteristics depending on whether the potential source is external (as in resistive heating) or internal (as in galvanic cells), and can approach reversible limits only in the latter case.

📌 Key points (3–5)

  • What electrical work is: energy transfer across a system boundary due to electric current flowing through a circuit, treated as work because it could equivalently be done by raising/lowering weights.
  • Two main scenarios: resistive heating (external voltage source, fully dissipative) versus galvanic cells (internal potential source, partially recoverable).
  • Common confusion: not all electrical energy transfer is "work"—if current doesn't cross the system boundary (e.g., redefining the system to exclude the resistor), the energy transfer becomes heat instead.
  • Reversible limits: galvanic cell work approaches a finite reversible limit as current approaches zero; resistive heating work is always dissipative with zero reversible work.
  • Minimal work principle: for irreversible processes with finite rates, work is least positive (or most negative) in the limit of infinite slowness.

⚡ Fundamentals of electrical work in circuits

⚡ Electric potential and charge flow

Electric potential φ at a point: the work needed to reversibly move an infinitesimal test charge from infinity to that point, divided by the test charge value.

  • Electrical potential energy of a charge = φ × charge.
  • In a circuit, electrons enter through one conductor (the "right" conductor at potential φ_R) and leave through another (the "left" conductor at potential φ_L).
  • Each electron carries charge −e (negative, where e is the elementary charge).
  • Energy difference for an electron between the two conductors = (φ_R − φ_L) × (−e).

🔌 General formula for electrical work

When infinitesimal charge ∂Q_sys enters at the right conductor and leaves at the left:

∂w_el = Δφ ∂Q_sys (Equation 3.8.1)

where Δφ = φ_R − φ_L is the electric potential difference.

  • ∂Q_sys is the inexact differential (path function) representing cumulative charge entering at the right conductor since the process began.
  • Alternative form using current I = ∂Q_sys / dt: ∂w_el = I Δφ dt (Equation 3.8.3).
  • Sign convention: ∂Q_sys is negative when electrons enter at right, positive when they leave; I follows the same sign convention.

🧱 Why treat it as work, not heat?

  • The energy transfer is consistent with the work interpretation: the only effect on surroundings could be a change in elevation of an external weight.
  • Example: a sinking weight drives a generator that does electrical work on the system; electrical work by the system runs a motor that raises the weight.
  • Equal numbers of electrons enter and leave, so the system can be treated as closed (constant number of electrons).
  • The net energy depends only on the potential difference, not on individual energies at each conductor (which include unmeasurable rest-mass energy of electrons).

Don't confuse: The terms φ_R ∂Q_sys and φ_L ∂Q_sys are not the actual energies transferred at each conductor individually—only their difference is measurable and meaningful.

🔥 Resistive heating (dissipative work)

🔥 Setup and mechanism

  • System: electrical resistor immersed in liquid (or resistor + liquid together).
  • External voltage source provides potential difference Δφ across wires.
  • When electrons flow, the resistor warms due to ohmic resistance—inelastic collisions between moving electrons and stationary atoms (a type of friction).
  • Thermal energy then transfers from resistor to liquid if temperature difference exists.

📐 Work formulas for resistive heating

Starting from general formulas and applying Ohm's law (R_el = Δφ / I):

  • ∂w_el = I R_el ∂Q_sys (Equation 3.8.4)
  • ∂w_el = I² R_el dt (Equation 3.8.5)
  • Integrated form (constant I and R_el): w_el = I R_el Q_sys

🚫 Characteristics of dissipative work

  • When the voltage source is in the surroundings, I and Q_sys have the same sign → w_el is always positive for finite current.
  • w_el = 0 only when I = 0 (no current).
  • Work is positive for both positive and negative changes of Q_sys.
  • Reversible limit is zero: as the rate (current I) approaches zero, work vanishes.
  • This is fully dissipative work—energy cannot be recovered by reversing the process.

Example: Figure 3.16 shows work versus current for fixed |Q_sys|—the curve passes through zero at I = 0, resembling the dissipative stirring work graph.

🔄 System boundary matters

  • If you redefine the system to be only the liquid (excluding the resistor):
    • Electric current no longer crosses the system boundary.
    • Energy transfer from resistor to liquid must be classified as heat, not work.

Don't confuse: Whether energy transfer is "work" or "heat" depends on how you draw the system boundary.

🔋 Galvanic cells (work with reversible limit)

🔋 What is a galvanic cell?

Galvanic cell: an electrochemical system that, when isolated, exhibits an electric potential difference between two terminals; the potential difference originates at interfaces between phases within the cell.

  • When current passes through the cell, a cell reaction occurs (chemical transformation).
  • The cell can do work on external components (e.g., a resistor) using energy from the cell reaction.

⚙️ Cell potential and internal resistance

Cell potential E_cell = φ_R − φ_L (Equation 3.8.6)

  • When isolated at zero current (equilibrium): E_cell = E_cell,eq (equilibrium cell potential).
  • When current flows: E_cell = E_cell,eq + I R_cell (Equation 3.8.7), where R_cell is internal resistance.
    • I < 0 (electrons entering at right) → E_cell < E_cell,eq
    • I > 0 (electrons leaving at right) → E_cell > E_cell,eq

🔄 Electrical work formula for galvanic cells

∂w_el = E_cell ∂Q_sys (Equation 3.8.8)

  • In the circuit shown (Figure 3.17b), E_cell is positive and ∂Q_sys is negative → ∂w_el is negative (cell does work on surroundings).
  • Work is irreversible at finite current due to internal resistance (energy dissipation).

🎯 Reversible limit

To approach reversible work:

  • Replace external resistor with an adjustable voltage source to control E_cell and I.
  • As I → 0 from either direction, E_cell → E_cell,eq.
  • Reversible work: ∂w_el,rev = E_cell,eq ∂Q_sys (Equation 3.8.9)

Key difference from resistive heating:

  • Electrical work is least positive or most negative in the reversible limit (nonzero).
  • Energy transferred can be partially recovered by returning Q_sys to its initial value.
  • The work coordinate Q_sys is proportional to the extent of the cell reaction, which is a state function (consistent with the principle that work with a reversible limit has a state-function work coordinate).

Example: Figure 3.18 shows work versus current—two branches approach finite (nonzero) reversible limits as I → 0, unlike the dissipative case.

🧮 Minimal work principle

🧮 General statement

For an irreversible adiabatic process of a closed system:

  • Starting and ending in equilibrium states.
  • Work coordinate X changes at a finite rate.
  • For a given initial state and given ΔX:
    • Work is less positive or more negative the more slowly X changes.
    • Work is least positive or most negative in the limit of infinite slowness.

📊 Mathematical formulation

SymbolMeaning
w_irrWork during irreversible process at finite rate
w_0Adiabatic work for same initial state and ΔX in the limit of infinite slowness
w_irr − w_0Always positive
  • If the work has a reversible limit, then w_0 = w_rev.
  • The minimal work principle applies to:
    • Expansion work (Section 3.5.5)
    • Work in gravitational fields (Section 3.6)
    • Electrical work with galvanic cells (Section 3.8.3)

🔍 Comparison of work types

Work typeReversible limitEnergy recoveryExample figure
Shaft work with frictionZeroNone (fully dissipative)Fig. 3.12(b)
Shaft work without frictionNonzeroPartialFig. 3.12(a)
Resistive heatingZeroNone (fully dissipative)Fig. 3.16
Galvanic cellNonzero (E_cell,eq ∂Q_sys)PartialFig. 3.18

Don't confuse: "Irreversible" doesn't always mean "fully dissipative"—galvanic cells have irreversible work at finite current but still approach a finite reversible limit, allowing partial energy recovery.

18

Irreversible Work and Internal Friction

3.9 Irreversible Work and Internal Friction

🧭 Overview

🧠 One-sentence thesis

Irreversible processes at finite rates require more work input (or yield less work output) than reversible processes with the same state change, and the difference can be attributed to internal friction that dissipates energy into thermal energy within the system.

📌 Key points (3–5)

  • Minimal work principle: For a given initial state and change in work coordinate, work is least positive (or most negative) in the limit of infinite slowness—the reversible limit.
  • Internal friction: The positive difference between irreversible work and reversible work (w_irr − w_0) can be attributed to internal friction dissipating other forms of energy into thermal energy.
  • Partial vs complete dissipation: Energy dissipation can be partial (e.g., lubricated friction in a piston) or complete (dissipative work like stirring or electrical heating with no reversible limit).
  • Common confusion: Not all irreversible processes involve internal friction—spontaneous heat flow across boundaries or homogeneous chemical reactions do not involve internal friction, nor does reversible adiabatic compression where thermal energy increases without friction.
  • Rate dependence: The faster the process, the more positive the work difference (w_irr − w_0) becomes due to friction effects.

🔄 The Minimal Work Principle

🔄 What the principle states

Minimal work principle: For an irreversible adiabatic process with a given initial state and a given change ΔX in work coordinate, the work is least positive or most negative in the limit of infinite slowness.

  • "Least positive" means the minimum work that must be done on the system.
  • "Most negative" means the maximum work that can be done by the system on the surroundings.
  • The slower the rate of change of X, the closer the work approaches this limit.

📚 Examples from earlier sections

The excerpt references three earlier illustrations:

  • Expansion work (Sec. 3.5.5): slower expansion yields more work output from the gas.
  • Gravitational field work (Sec. 3.6): slower lifting or lowering approaches reversible work.
  • Galvanic cell electrical work (Sec. 3.8.3): slower current draw yields more electrical work output.

🧮 Defining the work difference

  • Let w_irr = work during an irreversible adiabatic process at finite rate.
  • Let w_0 = adiabatic work for the same initial state and same ΔX in the limit of infinite slowness.
  • According to the minimal work principle: w_irr − w_0 is positive.
  • w_0 is the reversible work when a reversible limit exists.

🔥 Internal Friction and Energy Dissipation

🔥 What internal friction means

Internal friction: A conceptual attribution for the positive value of (w_irr − w_0), representing dissipation of other forms of energy into thermal energy within the system.

  • Internal friction is a mechanism that converts mechanical or other energy into heat inside the system.
  • It explains why irreversible processes require extra work input or yield less work output.

❌ When internal friction is NOT involved

The excerpt emphasizes that not all irreversible processes involve internal friction:

  • Spontaneous heat flow: When a temperature gradient causes heat to flow across the system boundary.
  • Irreversible chemical reactions: Spontaneous reactions in a homogeneous phase.
  • Reversible adiabatic compression: Even though thermal energy increases during infinitely slow adiabatic compression of a gas, internal friction is absent—the process is reversible.

Don't confuse: Increase in thermal energy does not automatically mean internal friction is present; reversible processes can also increase thermal energy.

🔀 Partial vs complete dissipation

TypeDescriptionReversible limit?
Partial dissipationSome energy is dissipated by friction, but a reversible limit existsYes (at infinite slowness)
Complete dissipation (dissipative work)All work is converted to thermal energy; no reversible limitNo

Dissipative work examples:

  • Stirring work (Sec. 3.7.1)
  • Electrical heating (Sec. 3.8.2)

For dissipative work: The final equilibrium state of an adiabatic process can also be reached by a path where positive heat replaces the dissipative work. This is a special case of the minimal work principle.

🛠️ Model: Cylinder-Piston with Sliding Friction

🛠️ The physical setup

The excerpt uses a gas-filled cylinder-and-piston device (Fig. 3.19) as a model for partial energy dissipation:

  • A rod is attached to the piston and slides through a bushing fixed inside the cylinder.
  • System boundary: includes the gas, the rod, and the bushing (but not the piston or cylinder wall).
  • The rod-bushing interface provides an obvious source of internal friction.

📐 Work calculation with friction

The work across the system boundary is:

  • w = − integral of F_sys dx from x_1 to x_2
  • where x is piston position and F_sys is the force exerted by the system on the surroundings.

Rewriting in terms of gas volume V (with dV = A_s dx, where A_s is cylinder cross-section area):

  • w = − integral of p_sys dV from V_1 to V_2
  • where p_sys = F_sys / A_s (total force per unit area exerted by the system).

Key insight: A plot of p_sys vs V is an indicator diagram, and work equals the negative of the area under the curve.

⚙️ Decomposing the system force

The force F_sys can be written as:

  • F_sys = p·A_s + F_fric
  • where p is the gas pressure and F_fric is the frictional force on the rod (with sign opposite to piston velocity).

Substituting into the work expression:

  • w = − integral of p dV − integral of (F_fric / A_s) dV
  • First term: work of expanding or compressing the gas.
  • Second term: frictional work w_fric = − integral of (F_fric / A_s) dV.

Frictional work is positive or zero and represents energy dissipated within the system by internal sliding friction.

🧴 Lubricated vs Dry Friction

🧴 Lubricated friction

When the rod-bushing contact is lubricated:

  • A fluid lubricant film separates the two solid surfaces.
  • Adjacent fluid layers move relative to one another, creating shear stress.
  • Frictional force depends on lubricant viscosity, film area, and rod velocity.
  • As rod velocity approaches zero, frictional force also approaches zero.
  • In the limit of infinite slowness: F_fric and w_fric vanish, and the process becomes reversible with w = − integral of p dV.

⚡ Effects at finite velocity

When the piston moves at an appreciable finite rate with lubricated friction:

  • w_fric is positive.
  • Expansion: irreversible work is less negative than reversible work for the same volume increase.
  • Compression: irreversible work is more positive than reversible work for the same volume decrease.
  • These effects are consistent with the minimal work principle.

🌀 Additional friction contributions

Besides the frictional force on the rod, piston velocity affects:

  • Gas force on piston: At large finite velocities, this effect tends to further decrease F_sys during expansion and increase it during compression—an additional contribution to internal friction.
  • Turbulent flow: If present within the system, this is also a contribution to internal friction.

📊 Indicator Diagrams and Final States

📊 Comparing irreversible and reversible paths

Figure 3.20 shows indicator diagrams for adiabatic expansion and compression with internal lubricated friction:

  • Solid curves: p_sys for irreversible processes at constant finite piston velocity.
  • Dashed curves: p_sys = p along a reversible adiabat (same initial state).
  • Open circles mark initial and final equilibrium states.

Areas under curves confirm:

  • Expansion: work is less negative along the irreversible path than the reversible path.
  • Compression: work is more positive along the irreversible path than the reversible path.

🌡️ Final state differences

Because of the work differences:

  • Final states of irreversible processes have greater internal energies and higher temperatures and pressures than final states of reversible processes with the same volume change.
  • This can be seen from the positions of final equilibrium state points on the indicator diagrams.

🔁 Equivalent heat path

The overall state change during irreversible expansion or compression can also be reached by:

  1. Reversible adiabatic volume change.
  2. Followed by positive heat at constant volume.

Since ΔU must be the same for both paths:

  • The required heat equals w_irr − w_rev.
  • This is NOT equal to the frictional work w_fric, because thermal energy released by frictional work increases gas pressure, making:
    • w_irr − w_rev less than w_fric for expansion.
    • w_irr − w_rev greater than w_fric for compression.

General limitation: There seems to be no general method to evaluate energy dissipated by internal friction, especially for irreversible processes with both work and heat.

📈 Rate dependence

Figure 3.21 shows adiabatic expansion work with internal lubricated friction as a function of average rate of volume change (for fixed magnitude of ΔV):

  • Open circles indicate the reversible limits (at zero rate).
  • As the rate increases, work becomes less negative for expansion and more positive for compression.
  • The faster the process, the further the work deviates from the reversible limit.
19

Reversible and Irreversible Processes: Generalities

3.10 Reversible and Irreversible Processes: Generalities

🧭 Overview

🧠 One-sentence thesis

Reversible processes represent idealized limits where work and heat are fully recoverable, while irreversible processes involve energy dissipation that makes the work required for a given change depend on the rate at which the process occurs.

📌 Key points (3–5)

  • Reversible work characteristics: work coefficients and coordinates are state functions, and energy transferred as work is fully recovered in the reverse process.
  • Irreversible work behavior: finite-rate processes involve partial or complete energy dissipation, which could alternatively be accomplished by adding positive heat.
  • Adiabatic work limits: for adiabatic processes with a reversible limit, work for a given volume change is least positive or most negative in the reversible limit.
  • Common confusion: lubricated vs dry friction—lubricated friction approaches zero at zero velocity (allowing reversibility in the limit), while dry friction remains finite even at zero velocity (preventing reversibility).
  • State variables needed: equilibrium states require one more independent variable than the number of independent reversible work coordinates.

🔄 Reversible process characteristics

⚙️ Work calculation and state functions

Infinitesimal work during a process: δw = Σᵢ Yᵢ dXᵢ, where Xᵢ is the work coordinate of kind i and Yᵢ is the conjugate work coefficient.

  • For reversible work, both work coefficients and work coordinates are state functions.
  • This means their values depend only on the current state, not on how the system reached that state.
  • Example: In reversible expansion, pressure (work coefficient) and volume (work coordinate) are both state functions.

🔁 Full energy recovery

  • Energy transferred as work in a reversible process is fully recovered as work of the opposite sign in the reverse process.
  • By the first law, heat is also fully recovered in the reverse process.
  • Example: Reversible adiabatic expansion from state A to state B can be exactly reversed by reversible adiabatic compression from B back to A, recovering all work.

⚡ Irreversible work and energy dissipation

🌡️ Partial dissipation effects

  • When work occurs irreversibly at a finite rate, there is partial or complete dissipation of energy.
  • The dissipation results in a change that could also be accomplished with positive heat.
  • Example: An increase of thermal energy within the system.

Key observation from indicator diagrams:

  • For adiabatic expansion: work is less negative along the irreversible path than along the reversible path.
  • For adiabatic compression: work is more positive along the irreversible path than along the reversible path.
  • Final states of irreversible processes have greater internal energies and higher temperatures and pressures than reversible processes with the same volume change.

🔥 Dissipative work

Dissipative work: positive irreversible work with complete energy dissipation.

  • The work coordinate for dissipative work is not a state function.
  • Examples include:
    • Stirring work
    • Electrical heating work
  • Don't confuse: dissipative work (complete dissipation) vs work with partial dissipation (like friction during expansion/compression).

📉 Rate dependence of work

For adiabatic processes with a reversible limit and a given initial state and volume change:

  • Work is least positive or most negative in the reversible limit.
  • As the rate of volume change increases, work deviates from the reversible limit.
  • For expansion: work becomes less negative (less work extracted).
  • For compression: work becomes more positive (more work required).

Important asymmetry:

  • Expansion work and compression work have opposite signs.
  • Only in the reversible limit do they have the same magnitude.

🔧 Friction mechanisms

🛢️ Lubricated friction

  • The frictional force approaches zero in the limit of zero piston velocity.
  • This allows the process to approach reversibility as the rate decreases.
  • Curves on indicator diagrams shift position as the average rate of volume change approaches zero.
  • Example: A piston with a lubricated rod and bushing—as velocity decreases, friction decreases, and the process approaches the reversible limit.

🪨 Dry friction

  • Due to microscopic roughness of contacting surfaces.
  • The frictional force does not approach zero in the limit of zero piston velocity.
  • Typically independent of contact area and sliding rate.
  • Curves on indicator diagrams change little as the rate approaches zero.

Critical difference:

  • In the limit of infinite slowness with dry friction:
    • Work is negative for expansion and positive for compression.
    • Expansion work is smaller in magnitude than compression work.
    • The expansion process cannot be reversed as a compression process, regardless of piston velocity.
    • These processes are therefore irreversible even at zero rate.

📊 State variables and system description

🔢 Equilibrium state variables

The number of independent variables needed to describe equilibrium states of a closed system:

  • Equals one greater than the number of independent work coordinates for reversible work.
  • Possible choices: each work coordinate plus either temperature or internal energy.

Example: For a simple gas with only expansion work (one work coordinate, volume):

  • Need two independent variables: volume and temperature (or volume and internal energy).

🌀 Nonequilibrium states

  • Require greater (often much greater) number of independent variables than equilibrium states.
  • This is because nonequilibrium states have spatial variations, gradients, and time-dependent properties.

⚠️ Special exceptions

  • If the system has internal adiabatic partitions allowing different phases to have different temperatures, the number of independent variables equals the number of work coordinates plus the number of independent temperatures.
  • Along special boundaries (e.g., the triple line of a pure substance), volume and temperature alone may not determine all phase amounts.

🔬 Energy accounting with friction

📐 Equivalent heat representation

The overall change of state during irreversible expansion or compression can be represented as:

  1. A reversible adiabatic volume change.
  2. Followed by positive heat at constant volume.

Since ΔU must be the same for both paths:

  • The required heat equals w_irr minus w_rev.
  • This is not the value of the frictional work itself.

🧮 Why the difference?

  • Thermal energy released by frictional work increases the gas pressure.
  • This makes (w_irr minus w_rev):
    • Less than w_fric for expansion.
    • Greater than w_fric for compression.
  • There seems to be no general method to evaluate energy dissipated by internal friction.
  • It would be even more difficult for irreversible processes with both work and heat.
20

Types of Processes

4.1 Types of Processes

🧭 Overview

🧠 One-sentence thesis

Expansion work in a gas system is calculated from the pressure exerted by the gas on the moving boundary and the volume change, and the correct expression depends on whether the process is reversible or irreversible.

📌 Key points (3–5)

  • General expansion work formula: work equals the integral of the pressure exerted by the gas on the piston (p_b) times the volume change.
  • Reversible limit: when piston motion is sufficiently slow, the gas has uniform pressure p throughout, and p_b can be replaced by p in the work calculation.
  • Common confusion: the correct pressure to use is p_b (pressure exerted by the gas on the piston), not p_ext (external pressure in the surroundings)—these are equal only in reversible processes or when the piston is frictionless and stationary in initial/final states.
  • Sign convention: expansion (positive dV) gives negative work (system does work on surroundings); compression (negative dV) gives positive work (surroundings do work on system).
  • Generalization: the expansion work formula applies to any isotropic phase undergoing slow deformation, not just cylinder-and-piston devices.

🔧 Pressure at the moving piston

🔧 Pressure difference between piston and wall

The excerpt derives the pressure p_b exerted by the gas on the moving piston and compares it to the pressure p at the stationary cylinder wall.

p_b: the pressure exerted by the gas on the moving piston during expansion or compression.

  • The derivation uses kinetic-molecular theory, tracking a single gas molecule colliding with the moving piston.
  • The piston moves with velocity u; the molecule has initial x-velocity v_x,1.
  • After collision, the molecule's x-velocity changes to v_x,2 = -v_x,1 + 2u.
  • The time-averaged force exerted by one molecule on the piston is calculated, then summed over all molecules and divided by piston area to get p_b.

📐 Relationship between p_b and p

The excerpt gives the formula:

p_b = p × [1 - 2u⟨v_x,1⟩ / ⟨v_x,1²⟩]

  • When the piston is stationary (u = 0), p_b equals p.
  • For a moving piston, p_b differs from p by a factor that depends on piston speed u and molecular velocity averages.
  • Example: nitrogen gas at 300 K with piston moving at 10 m/s → p_b is only about 5% lower than p during expansion, 5% higher during compression.
  • At low piston speeds, the percentage difference is proportional to piston speed, so for reasonably slow speeds the difference is quite small and can usually be ignored.

Don't confuse: p_b (pressure at the moving piston) vs. p (pressure at the stationary wall)—they are approximately equal only when the piston moves slowly.

🧮 Expansion work formulas

🧮 General formula for expansion work

The excerpt defines the system as the gas inside the cylinder. The only moving portion of the boundary is the inner surface of the piston.

Starting from the differential work formula ∂w = F_gas dx_pis, where F_gas is the force exerted by the system on the surroundings:

  • Substitute F_gas = p_b A_s (from the force-pressure relation).
  • Change the work coordinate from x_pis to V using dV = A_s dx_pis.

Result:

∂w = p_b dV (expansion work, closed system)

For a finite volume change:

w = - ∫(V1 to V2) p_b dV (expansion work, closed system)

  • The integral is a line integral; to evaluate it, p_b must be expressed as a function of V along the process path.
  • The negative sign in the integral ensures the correct sign convention.

⚖️ Reversible expansion work

When piston motion is sufficiently slow:

  • The gas has uniform pressure p throughout.
  • Work can be calculated as if the process has reached its reversible limit.
  • p_b can be replaced by p in the work formulas.

∂w = p dV (reversible expansion work, closed system)

w = - ∫(V1 to V2) p dV (reversible expansion work, closed system)

Important rule: An equation containing the symbol of an intensive property (like p) not assigned to a specific phase is valid only if that property is uniform throughout the system.

📏 Dimensional analysis tip

The product of pressure and volume has dimensions of energy:

  • 1 Pa·m³ = 1 J

This is helpful when checking units in expansion work calculations.

🔀 Sign convention and direction

🔀 Expansion vs compression

ProcessVolume changeSign of ∂wWork direction
ExpansionPositive dVNegativeSystem does work on surroundings
CompressionNegative dVPositiveSurroundings do work on system

Example: During expansion, the gas pushes the piston outward, doing work on the surroundings, so w is negative.

Example: During compression, the surroundings push the piston inward, doing work on the gas, so w is positive.

⚠️ Common confusion: p_b vs p_ext

⚠️ Which pressure to use

The excerpt emphasizes a common error in textbooks:

  • Incorrect: Some texts state that expansion work should be calculated from w = ∫ p_ext dV, where p_ext is a pressure in the surroundings.
  • Correct: The general expression is w = - ∫ p_b dV, where p_b is the pressure exerted by the gas on the piston.

🔍 Why p_b is correct

The reasoning:

  • The system is the gas.
  • The force exerted by the system on the surroundings is F_gas = p_b A_s.
  • The force F_ext = p_ext A_s is exerted by one part of the surroundings on another part of the surroundings.
  • Work is defined by the force exerted by the system on the surroundings, so w = - ∫ F_gas dx_pis = - ∫ p_b dV.

🔄 When p_b equals p_ext

The integrals ∫ F_gas dx_pis and ∫ F_ext dx_pis are equal in two cases:

  1. Reversible process: In the limit of infinite slowness, the piston has no friction (F_fric = 0) and no acceleration (F_net = 0), so F_gas = F_ext throughout the process.
  2. Frictionless piston, stationary in initial and final states: Both F_fric and ∫ F_net dx_pis are zero. (The integral ∫ F_net dx_pis equals the change in kinetic energy of the piston.)

Don't confuse: In the general irreversible case, ∫ F_gas dx_pis and ∫ F_ext dx_pis are NOT equal, so using p_ext instead of p_b gives the wrong answer.

🌐 Generalization to isotropic phases

🌐 Beyond cylinder-and-piston

Expansion work does not require a cylinder-and-piston device. The formula applies to any isotropic fluid or solid phase undergoing deformation.

Setup:

  • The system is an isotropic phase of arbitrary shape.
  • Various portions of the boundary undergo displacements in different directions.
  • The deformation is carried out slowly, so the pressure p remains uniform throughout the phase.

🧩 Work at a surface element

Consider a small surface element of the boundary with area A_s,σ:

  • Because the phase is isotropic, the force F_sys,σ = p A_s,σ exerted by the system on the surroundings is perpendicular to the surface element (no shearing force).
  • For an infinitesimal displacement ds_σ that reduces the volume, the volume change is dV_σ = A_s,σ ds_σ.
  • The work at this surface element is ∂w_σ = p dV_σ.

By summing the work over the entire boundary, the total reversible expansion work is the same integral form as before.

Example: Compression of a fluid blob of arbitrary shape—each small surface element contributes work according to the local displacement and area, but the total work still follows w = - ∫ p dV.

21

Statements of the Second Law

4.2 Statements of the Second Law

🧭 Overview

🧠 One-sentence thesis

The excerpt does not contain statements of the Second Law of thermodynamics; instead, it discusses expansion work, reversible processes, and applications to ideal gases under the First Law.

📌 Key points (3–5)

  • Content mismatch: The provided text covers expansion work and ideal gas processes (Chapter 3, First Law), not the Second Law as indicated by the title.
  • Main topics present: Reversible expansion work, work integrals for piston systems, and ideal gas behavior during isothermal and adiabatic processes.
  • Key formulas discussed: Work expressions like w = p dV for reversible processes, and temperature-volume relations for adiabatic expansions.
  • Common confusion: Reversible vs irreversible work—the integrals of F_gas and F_ext are equal only under specific conditions (infinite slowness or frictionless stationary states).

⚠️ Content note

⚠️ Excerpt does not match title

The title "4.2 Statements of the Second Law" suggests content about entropy, spontaneity, or different formulations of the Second Law (e.g., Clausius, Kelvin-Planck statements). However, the provided text is from Chapter 3 on the First Law and covers:

  • Expansion work mechanics (Section 3.4)
  • Applications to ideal gases (Section 3.5)
  • Isothermal and adiabatic processes
  • Indicator diagrams

The excerpt contains no discussion of the Second Law, its statements, or related concepts like entropy or irreversibility in the thermodynamic sense.

🔧 Expansion work concepts (from excerpt)

🔧 When work integrals are equal

The excerpt explains that the integrals of F_gas and F_ext over piston displacement are equal in two specific cases:

  • Infinite slowness: No friction (F_fric = 0) and no acceleration (F_net = 0), so gas force equals external force throughout.
  • Frictionless with stationary endpoints: Both friction and the integral of net force times displacement are zero (the latter equals the change in piston kinetic energy).

In general irreversible cases, these integrals are not equal.

🌐 Work for isotropic phases

Expansion work does not require a cylinder-and-piston device.

  • For an isotropic fluid or solid phase with uniform pressure p, the work expression ∂w = p dV applies regardless of boundary shape.
  • The force exerted by the system is perpendicular to each surface element (no shearing force).
  • By summing work over the entire boundary, the total reversible expansion work has the same form as for a piston-cylinder system.
  • Not valid for anisotropic materials: The angle in the work equation might not be zero.

📐 Work coordinates rule

The number of independent variables needed to describe equilibrium states of a closed system is one greater than the number of independent work coordinates for reversible work.

  • For a system with reversible work of the form ∂w_i = Y_i dX_i, the number of independent variables equals (number of work coordinates + 1).
  • Example: For expansion work (one coordinate, V), two independent variables (e.g., p and V, or T and V) describe equilibrium states.

🎈 Ideal gas applications (from excerpt)

🎈 Ideal gas definition

The excerpt provides a two-part definition:

  1. Equation of state: pV = nRT
  2. Internal energy: U depends only on temperature in a closed system

Molecular basis: A gas with negligible intermolecular interactions fulfills both requirements. Real gases approach ideal behavior at large molar volumes (low pressure), where intermolecular distances are large.

🌡️ Isothermal expansion (constant temperature)

For reversible isothermal expansion or compression of an ideal gas:

  • Work = nRT times the natural logarithm of (V_final / V_initial)
  • Derived by substituting p = nRT/V into the work integral and treating n and T as constants.

❄️ Adiabatic expansion (no heat transfer)

Key relations for reversible adiabatic processes:

QuantityRelationNotes
Internal energy changedU = C_V dTValid for any process of closed ideal gas system
Temperature-volumeT_final = T_initial × (V_initial / V_final)^(nR/C_V)Temperature decreases during expansion
Pressure-volumep_final = p_initial × (V_initial / V_final)^(1 + nR/C_V)Adiabat crosses isotherms of lower T during expansion
Workw = C_V × (T_final - T_initial)Equals change in internal energy

Why temperature changes: During adiabatic expansion, the gas does work on surroundings without compensating heat flow, so internal energy (and thus temperature) decreases.

📊 Indicator diagrams

An indicator diagram (or pressure–volume diagram) is usually a plot of p as a function of V.

  • The area under the curve equals the integral of p dV, which is the negative of reversible expansion work.
  • Direction matters: rightward (expansion) gives positive area and negative work; leftward (compression) gives negative area and positive work.
  • Example: The adiabat (solid curve) shows pressure vs volume during reversible adiabatic process; it crosses isotherms (dashed curves) of progressively lower temperatures during expansion.
22

4.3 Concepts Developed with Carnot Engines

4.3 Concepts Developed with Carnot Engines

🧭 Overview

🧠 One-sentence thesis

The excerpt develops foundational thermodynamic concepts through reversible and spontaneous processes, showing that reversible adiabatic processes represent limiting cases where the system does maximum work on the surroundings or the surroundings do minimum work on the system.

📌 Key points (3–5)

  • Reversible adiabatic expansion: follows a specific pressure-volume relationship (adiabat) and crosses isotherms of progressively lower temperatures as the gas cools while doing work without heat input.
  • Indicator diagrams: plots of pressure vs. volume where the area under the curve equals the magnitude of reversible expansion work; historically used to monitor steam engine performance.
  • Spontaneous vs. reversible work: for the same initial state and volume change, spontaneous adiabatic processes involve less negative work (expansion) or more positive work (compression) than reversible processes.
  • Common confusion: pressure at a moving boundary (p_b) differs from pressure at stationary walls during rapid spontaneous processes; reversible processes assume uniform pressure throughout.
  • Free expansion: expansion into a vacuum involves no work and no temperature change for an ideal gas, because there is no displacement at contact forces.

🌡️ Reversible adiabatic processes

🌡️ Pressure-volume relationship

Adiabat: the curve showing how pressure varies with volume during a reversible adiabatic expansion or compression.

  • The relationship is given by: p₂ = p₁ × (V₁/V₂) raised to the power (1 + nR/C_V), where n is amount of substance, R is gas constant, and C_V is heat capacity at constant volume.
  • This applies specifically to ideal gases undergoing reversible adiabatic processes.
  • The solid curve (adiabat) is distinct from isotherms (dashed curves showing constant-temperature behavior).

❄️ Temperature behavior during expansion

  • In the direction of increasing volume (expansion), the adiabat crosses isotherms of progressively lower temperatures.
  • Why cooling occurs: the gas loses energy by doing work on the surroundings without any compensating heat flow into the system.
  • This is a direct consequence of the first law: internal energy decreases because work is done (negative) and heat is zero (adiabatic).
  • Example: as an ideal gas expands adiabatically and reversibly, it simultaneously cools because energy leaves as work.

📊 Indicator diagrams and work visualization

📊 What an indicator diagram shows

Indicator diagram (or pressure–volume diagram): usually a plot of pressure p as a function of volume V, describing the path of an expansion or compression process of a fluid that is essentially uniform.

  • The curve describes the process path for a fluid.
  • The area under the curve equals the integral of p dV, which is the negative of the reversible expansion work (w = -∫p dV).
  • Direction matters:
    • Expansion (rightward): area is positive, work is negative (system does work on surroundings).
    • Compression (leftward): area is taken as negative, work is positive (surroundings do work on system).

📊 Generalized definition

  • More generally, an indicator diagram can plot any work coefficient (or its negative) as a function of the work coordinate.
  • Example: plot pressure p_b at a moving boundary as a function of V; the area equals ∫p_b dV, the negative of expansion work in general.

🕰️ Historical context

  • Invented by James Watt in the late 1700s to monitor steam engine performance.
  • Original design: a simple pressure gauge with a piston in a small secondary cylinder, opposed by a compressed spring.
  • Later versions: a pencil attached to the piston traced a closed curve on a paper-covered roll linked to the main cylinder piston.
  • The area of the closed curve was proportional to the net work performed by one engine cycle.

⚡ Spontaneous vs. reversible processes

⚡ Pressure differences during rapid processes

  • During rapid spontaneous expansion: pressure p_b at the moving piston is less than the pressure at the stationary wall.
  • During rapid spontaneous compression: pressure p_b at the moving piston is greater than the pressure at the stationary wall.
  • Don't confuse: in reversible processes, the system is assumed to be uniform with the same pressure throughout; spontaneous processes involve pressure gradients.

⚡ Work comparison for adiabatic processes

Process typeExpansion workCompression work
Reversible adiabaticMost negative (maximum work done by system)Least positive (minimum work done on system)
Spontaneous adiabaticLess negative than reversibleMore positive than reversible
  • For the same initial state and same volume change:
    • Spontaneous adiabatic expansion: work is less negative than reversible (w = -∫p_b dV with smaller p_b).
    • Spontaneous adiabatic compression: work is more positive than reversible (w = -∫p_b dV with larger p_b).

⚡ General principle for adiabatic processes

  • For an adiabatic expansion or compression with a given change of the work coordinate, starting at a given initial equilibrium state, the work is algebraically smallest (least positive or most negative) in the reversible limit.
  • In other words:
    • Surroundings do the least possible work on the system in the reversible limit.
    • System does the maximum possible work on the surroundings in the reversible limit.
  • This behavior applies to any adiabatic process of a closed system.

🌌 Free expansion into a vacuum

🌌 What happens during free expansion

Free expansion: the process when gas expands from one vessel into an evacuated vessel after opening a stopcock.

  • The system is the gas.
  • Surroundings exert contact force only at the vessel walls, where there is no displacement.
  • Therefore: no work in free expansion (∂w = 0).

🌌 Energy and temperature effects

  • If carried out adiabatically (thermally insulated apparatus): no heat and no work, so no change in internal energy (ΔU = 0).
  • For an ideal gas: internal energy depends only on temperature.
  • Conclusion: adiabatic free expansion of an ideal gas causes no temperature change.
  • Example: opening a valve between a gas-filled chamber and a vacuum chamber in an insulated container leaves the gas temperature unchanged (for ideal gas behavior).

🪐 Work in a gravitational field

🪐 Setup and force components

  • A spherical body (e.g., glass marble) immersed in a liquid or gas in an external gravitational field.
  • Elevation z is measured above the bottom of the vessel; displacements are parallel to the vertical z axis.
  • Work is given by: ∂w = F_sur,z dz, where F_sur,z is the upward component of the net contact force exerted by surroundings on the system at the moving boundary.
  • The gravitational force does not contribute to F_sur,z (as explained in an earlier section).

🪐 Free fall through fluid

System choice 1: body + fluid combined

  • System boundary is where fluid contacts atmosphere and vessel walls.
  • No displacement of this boundary → no work (∂w = 0).
  • If adiabatic: internal energy remains constant; as the body loses gravitational potential energy, the system gains equal kinetic and thermal energy.

System choice 2: body only (fluid in surroundings)

  • Upward forces on the body:
    • Buoyant force: F_buoy = ρV₀g (where ρ is fluid density, V₀ is body volume, g is acceleration of free fall).
    • Frictional drag force: F_fric (opposite sign from velocity v = dz/dt).
    • Gravitational force -mg is not included in F_sur,z.
  • Gravitational work: ∂w = F_sur,z dz = (F_buoy + F_fric) dz.
  • Work is negative because dz is negative (body falls): the body does work on the fluid.
  • The quantity |F_buoy dz| is work of moving displaced fluid upward; |F_fric dz| is energy dissipated by friction to thermal energy in surroundings.
  • No reversible limit for free fall: the rate of energy transfer cannot be controlled from the surroundings and cannot approach zero.

🪐 Body suspended by string

System choice 1: body + fluid (string in surroundings)

  • Moving boundary is at the point where string attaches to body.
  • String exerts upward force F_str.
  • Gravitational work: ∂w = F_sur,z dz = F_str dz.
  • Newton's second law: (-mg + F_buoy + F_fric + F_str) = m dv/dt.
  • Solving for F_str: F_str = -mg - F_buoy - F_fric + m dv/dt.
  • Work expression: ∂w = F_str dz = (-mg - F_buoy - F_fric + m dv/dt) dz.
  • Work can be positive or negative depending on whether the body is pulled up or lowered.
  • The term (m dv/dt) dz is an infinitesimal change of the body's kinetic energy E_k.

🪐 Interpreting the work terms

  • The work includes contributions from:
    • Gravitational potential energy change (related to -mg dz).
    • Buoyancy work (moving displaced fluid, -F_buoy dz).
    • Frictional dissipation (-F_fric dz).
    • Kinetic energy change (m dv/dt dz).
  • Don't confuse: the gravitational force itself does not appear in F_sur,z; only contact forces from surroundings (buoyancy, friction, string tension) contribute to the work term.
23

The Second Law for Reversible Processes

4.4 The Second Law for Reversible Processes

🧭 Overview

🧠 One-sentence thesis

The Clausius inequality, derived from the Kelvin–Planck statement, proves that entropy exists as a state function whose change in any reversible process equals the integral of heat divided by boundary temperature.

📌 Key points (3–5)

  • The Clausius inequality: For any cyclic process of a closed system, the cyclic integral of (heat/boundary temperature) cannot be positive; it equals zero only for reversible cycles.
  • Adiabatic inaccessibility principle: Every equilibrium state has other equilibrium states infinitesimally close that cannot be reached by reversible adiabatic processes—this leads to the existence of reversible adiabatic surfaces.
  • Entropy as a state function: Reversible adiabatic surfaces are surfaces of constant entropy; entropy change between any two states is path-independent when calculated via reversible processes.
  • Common confusion: The integral of (¶q/T_b) depends on the path for irreversible processes, but for reversible processes between the same two states it always gives the same entropy change regardless of path.
  • Extensivity: Entropy is an extensive property—the entropy of a system equals the sum of the entropies of its subsystems.

🔧 Deriving the Clausius inequality

🔧 The experimental setup

The excerpt constructs a "supersystem" to apply the second law:

  • An experimental system (arbitrary complexity, may have gradients, reactions, phase changes, etc.) undergoes a cyclic process.
  • A hypothetical Carnot engine exchanges heat with both the experimental system and a single heat reservoir at constant temperature T_res.
  • The supersystem (experimental system + Carnot engine) is closed and exchanges heat only with the single reservoir.
  • Work is controlled from outside the supersystem.

Why this setup matters: By restricting heat exchange to a single reservoir, the Kelvin–Planck statement (no cycle can convert heat from a single reservoir completely into work) can be applied directly.

🔧 Heat transfer in infinitesimal Carnot cycles

During each stage of the experimental process:

  • The Carnot engine undergoes many infinitesimal Carnot cycles.
  • In one isothermal step: the engine at temperature T_res exchanges heat ¶q′ with the reservoir.
  • In the other isothermal step: the engine at temperature T_b (matching the experimental system's boundary temperature) exchanges heat ¶q with the experimental system.
  • The relation from Carnot cycle properties: ¶q′ = T_res × (¶q / T_b).

After the experimental system completes its cycle, integrating gives the net heat entering the supersystem:

q′ = T_res × (cyclic integral of ¶q / T_b)

🔧 Applying the Kelvin–Planck statement

  • If q′ were positive, the supersystem would have converted heat from a single reservoir completely into work (since the supersystem returns to its initial state, internal energy change is zero, so net work would be negative).
  • The Kelvin–Planck statement says this is impossible.
  • Therefore q′ cannot be positive, which yields:

Clausius inequality: (cyclic integral of ¶q / T_b) ≤ 0 for any cyclic process of a closed system.

Valid only if: the cycle contains only reversible and irreversible changes, not impossible changes (like the reverse of an irreversible process).

Special case: If the entire cycle is adiabatic (only possible if reversible), the cyclic integral equals zero.

🌐 Reversible adiabatic surfaces and entropy existence

🌐 One-way heat and adiabatic inaccessibility

Consider a reversible process from state A to state B with "one-way heat" (all infinitesimal heat quantities have the same sign, either all positive or all zero; or all negative or all zero):

  • If heat flows into the system (positive one-way heat), the integral of (¶q_rev / T_b) from A to B is positive.
  • The Clausius inequality then requires that the integral from B back to A must be negative.
  • Conclusion: There is no adiabatic path from B to A (since adiabatic means ¶q = 0, so the integral would be zero, not negative).

General rule: Whenever state A can be changed to state B by a reversible process with finite one-way heat, it is impossible to change from either state to the other by a reversible adiabatic process.

Example: Heating a liquid reversibly increases its temperature; you cannot duplicate this temperature increase by reversible work alone (though you can do it irreversibly, e.g., by stirring).

Carathéodory's principle of adiabatic inaccessibility: Every equilibrium state has other equilibrium states infinitesimally close that are inaccessible by reversible adiabatic processes.

🌐 Reversible adiabatic surfaces

Visualize an N-dimensional state space where axes represent temperature T and each work coordinate:

  • A reversible adiabatic process (adiabatic boundary, slowly varying work coordinates) can connect any combination of work coordinate values.
  • Key question: Do all reversible adiabatic paths starting from a common initial state fill a volume or lie on a surface?
  • If they filled a volume, every nearby point would be accessible by reversible adiabatic process—contradicting adiabatic inaccessibility.
  • Answer: All reversible adiabatic paths from a common initial state lie on a unique (N−1)-dimensional surface in N-dimensional space.

Result: An infinite number of non-intersecting reversible adiabatic surfaces fill the state space.

  • A reversible process with positive one-way heat moves the state from one surface to another on one side.
  • A reversible process with negative one-way heat moves to a surface on the opposite side.

🌐 Defining entropy

Entropy S: A state function assigned the same value everywhere on one reversible adiabatic surface, and a different unique value on each different surface.

  • Reversible adiabatic surfaces are surfaces of constant entropy.
  • Since the surfaces fill the space without intersecting, every equilibrium state lies on exactly one surface with a definite value of S.
  • Ordering: Reversible processes with positive one-way heat correspond to increasing S; negative one-way heat corresponds to decreasing S.

Don't confuse: The existence of entropy is proven by the existence of reversible adiabatic surfaces; the actual numerical values assigned are somewhat arbitrary (one reference state is chosen, and differences are calculated).

📐 Formula for entropy change

📐 Path independence for reversible processes

Consider two different reversible paths (path 1: A→B, path 2: C→D) connecting the same pair of reversible adiabatic surfaces:

  • Combine the experimental system with a Carnot engine and heat reservoir as before.
  • Net heat entering supersystem along path 1: q′ = T_res × (integral of ¶q_rev / T_b from A to B).
  • Net heat along path 2: q″ = T_res × (integral of ¶q_rev / T_b from C to D).
  • Construct a cycle: A→B→D→C→A (where B→D and C→A are reversible adiabatic).
  • Net heat in this cycle: q′ − q″.
  • The reverse cycle has net heat q″ − q′.
  • By Kelvin–Planck, neither cycle can have positive net heat, so q′ = q″.

Conclusion: The integral of (¶q_rev / T_b) has the same value for any reversible path between the same two reversible adiabatic surfaces (i.e., between states with the same entropy difference).

📐 Entropy change formula

Entropy change for a reversible process in a closed system:
ΔS = integral of (¶q_rev / T_b)

For an infinitesimal reversible path element:

dS = ¶q_rev / T_b

Why this works:

  • Makes ΔS positive when heat flows in (positive one-way heat).
  • Makes ΔS negative when heat flows out (negative one-way heat).
  • Makes ΔS zero for adiabatic processes.
  • Gives the same ΔS for any reversible path between the same two states.
  • Makes ΔS additive for consecutive processes.

Integrating factor: The quantity 1/T_b is an integrating factor that converts ¶q_rev (not an exact differential) into dS (an exact differential of a state function).

📐 Alternative derivation via cyclic integral

A more direct proof that entropy is a state function:

  • For a reversible process, reversing the path changes the sign but not magnitude of the integral of (¶q_rev / T_b).
  • For a reversible cycle, the Clausius inequality says the cyclic integral cannot be positive.
  • If it were negative, the reverse cycle would have a positive cyclic integral—also impossible.
  • Therefore: (cyclic integral of ¶q_rev / T_b) = 0 for any reversible cycle.
  • This means the integral between any two states A and B is path-independent (same value along any reversible path).
  • Path independence makes (¶q_rev / T_b) an exact differential, so entropy is a state function.

Example: Choose any two reversible paths from A to B; the entropy change S_B − S_A calculated via either path is identical.

🧱 Properties of entropy

🧱 Extensivity

Entropy is an extensive property:

  • Divide a system at uniform temperature T into two subsystems A and B.
  • For a reversible infinitesimal change: dS_A = ¶q_A / T, dS_B = ¶q_B / T, dS = ¶q_rev / T.
  • Since ¶q_rev = ¶q_A + ¶q_B, we have dS = dS_A + dS_B.
  • Entropy changes are additive, so S = S_A + S_B.

🧱 Evaluating entropy of a state

For an equilibrium state:

  • Assign an arbitrary value to one reference state.
  • Calculate ΔS along a reversible path from the reference state to the state of interest using ΔS = integral of (¶q_rev / T_b).

For a nonequilibrium state:

  • Imagine imposing hypothetical internal constraints (e.g., rigid adiabatic partitions) that convert the nonequilibrium state into a constrained equilibrium state with the same internal structure.
  • Evaluate the entropy of this constrained equilibrium state.

Don't confuse: Entropy is defined for equilibrium states via reversible processes; for nonequilibrium states, a hypothetical constrained equilibrium is used.

24

The Second Law for Irreversible Processes

4.5 The Second Law for Irreversible Processes

🧭 Overview

🧠 One-sentence thesis

For irreversible processes in a closed system, the entropy change is always greater than the heat transferred divided by boundary temperature, replacing the equality that holds for reversible processes with an inequality.

📌 Key points (3–5)

  • Reversible vs irreversible: For reversible processes, dS equals ∂q/T_b; for irreversible processes, dS is greater than ∂q/T_b (equalities become inequalities).
  • Irreversible adiabatic processes: No reversible adiabatic path can connect the initial and final states of an irreversible adiabatic process; the entropy change must be positive.
  • Common confusion: Nonequilibrium states need special handling—imagine imposing hypothetical internal constraints to create a constrained equilibrium state with the same entropy.
  • General irreversible processes: The derivation extends from adiabatic cases to all irreversible processes using a supersystem with a Carnot engine and heat reservoir.
  • Why it matters: This inequality establishes that irreversible processes always increase entropy, a fundamental constraint on real-world thermodynamic changes.

🔧 Entropy as a state function

🔧 How entropy is defined for any state

Entropy S is a state function: its value depends only on the current state, not on the system's history or future.

  • The excerpt shows that the entropy change ΔS between states A and B is calculated by integrating ∂q_rev/T_b along any reversible path connecting them.
  • Because this integral depends only on the endpoints A and B, entropy itself is a state function.
  • To find the entropy of a particular equilibrium state: assign an arbitrary value to one reference state, then evaluate the entropy change along a reversible path to the state of interest.

📏 Entropy is extensive

  • The excerpt proves that entropy is an extensive property (scales with system size).
  • Proof sketch: divide a uniform-temperature system into two closed subsystems A and B; the entropy changes are additive: dS = dS_A + dS_B, so S = S_A + S_B.
  • Example: if subsystem A has entropy 10 units and subsystem B has entropy 15 units, the combined system has entropy 25 units.

🧩 Entropy of nonequilibrium states

The excerpt addresses a conceptual challenge: how to assign entropy to a nonequilibrium state.

Procedure:

  • Imagine imposing hypothetical internal constraints that convert the nonequilibrium state into a constrained equilibrium state with the same internal structure.
  • Examples of such constraints: rigid adiabatic partitions between phases of different temperature/pressure, semipermeable membranes to prevent species transfer, inhibitors to prevent chemical reactions.
  • Assume these constraints can be imposed or removed reversibly without heat, so there is no entropy change.
  • If the system has macroscopic internal motion, imposing constraints involves negative reversible work to bring moving regions to rest.
  • If the system is nonuniform, the constraints partition it into practically-uniform regions whose entropy is additive.
  • The entropy of the nonequilibrium state equals the entropy of this constrained equilibrium state, found by integrating ∂q_rev/T_b along a reversible path from a known reference state.

Don't confuse: The entropy of a state with macroscopic internal motion is defined to be the same as the entropy of the same state "frozen in time" (without motion); purely mechanical processes have ΔS = 0.

🔥 Irreversible adiabatic processes

🔥 No reversible adiabatic path exists

The excerpt establishes a key result: for an irreversible adiabatic process from state A to state B, there is no reversible adiabatic path connecting A and B.

Case 1: Work occurs during the process (∂w ≠ 0):

  • Experience shows that the work would differ if the work coordinates changed at a different rate, because energy dissipation from internal friction would differ.
  • In the limit of infinite slowness, the process would become reversible, but the net work and final internal energy would differ from those of the irreversible process.
  • Therefore, the final state of the reversible adiabatic process is different from B—no reversible adiabatic path exists between A and B.

Case 2: No work occurs (∂w = 0):

  • The process occurs at constant internal energy U in an isolated system.
  • A reversible limit cannot be reached without heat or work.
  • Any reversible adiabatic change from state A would require work, causing a change in U and preventing the system from reaching state B.
  • Again, no reversible adiabatic path exists between A and B.

Implication: The only reversible paths between A and B must be nonadiabatic (involve heat transfer).

➕ Entropy change must be positive

Since no reversible adiabatic path exists, the entropy change ΔS_A→B (evaluated by integrating ∂q_rev/T_b over a reversible path) cannot be zero.

Could ΔS_A→B be negative?

  • Along the irreversible adiabatic process A→B, ∂q = 0, so the integral of ∂q/T_b is zero.
  • Imagine completing a cycle by returning along a different, reversible path from B back to A.
  • The Clausius inequality says the integral of ∂q_rev/T_b along the reversible return path cannot be positive.
  • But this integral equals −ΔS_A→B, so ΔS_A→B cannot be negative.

Conclusion: ΔS_A→B cannot be zero and cannot be negative, so it must be positive.

For infinitesimal changes: If states A and B are infinitesimally different, then for an infinitesimal irreversible adiabatic change, dS must be positive.

🌐 General irreversible processes

🌐 Supersystem construction

To extend the result to nonadiabatic irreversible processes, the excerpt uses a clever construction involving a supersystem.

Setup:

  • The supersystem includes: the experimental system, a Carnot engine, and a heat reservoir of constant temperature T_res.
  • During an irreversible change of the experimental system, the Carnot engine undergoes many infinitesimal cycles.
  • In each cycle, the Carnot engine exchanges heat ∂q′ at temperature T_res with the heat reservoir and heat ∂q at temperature T_b with the experimental system.
  • Sign convention: ∂q′ is positive if heat is transferred to the Carnot engine; ∂q is positive if heat is transferred to the experimental system.

Key feature: The supersystem exchanges work, but not heat, with its surroundings—it is adiabatic overall.

🔢 Deriving the inequality

Entropy changes during one infinitesimal cycle:

  • Carnot engine: net entropy change is zero (it completes a cycle).
  • Experimental system: entropy change is dS.
  • Heat reservoir: receives heat ∂q′ = T_res · (∂q / T_b) from the Carnot engine, so its entropy change is dS_res = −∂q′ / T_res = −∂q / T_b.

Supersystem entropy change:

  • dS_ss = dS + dS_res = dS − ∂q / T_b.

Applying the adiabatic result:

  • The process within the supersystem is adiabatic and includes an irreversible change within the experimental system.
  • From the earlier result (Sec. 4.5.1), dS_ss must be positive.
  • Therefore: dS − ∂q / T_b > 0, which means dS > ∂q / T_b.

⚖️ Reversible vs irreversible comparison

Process typeEntropy changeRelation to heat transfer
ReversibledS = ∂q / T_bEquality holds
IrreversibledS > ∂q / T_bInequality: entropy change is greater

Don't confuse: The inequality does not say that irreversible processes produce more heat; it says that for the same heat transfer ∂q, the entropy change dS is larger in an irreversible process.

Example: If an experimental system receives the same amount of heat ∂q in two different processes—one reversible, one irreversible—the irreversible process will have a larger entropy increase dS.

25

Applications of the Second Law

4.6 Applications

🧭 Overview

🧠 One-sentence thesis

The mathematical statement of the second law—combining the equality dS = ∂q_rev/T_b for reversible changes and the inequality dS > ∂q/T_b for irreversible changes—provides a unified framework for calculating entropy changes in heating, expansion, spontaneous processes, and isolated systems, with the key insight that entropy continuously increases during spontaneous processes until equilibrium is reached.

📌 Key points (3–5)

  • Unified mathematical statement: The second law combines an equality (reversible: dS = ∂q_rev/T_b) and an inequality (irreversible: dS > ∂q/T_b) into one relation dS ≥ ∂q/T_b.
  • Entropy always increases in isolated systems: During any spontaneous, irreversible process in an isolated system, entropy continuously increases until it reaches a maximum at equilibrium.
  • Reversible heating and expansion: Heating increases entropy (ΔS = C ln(T₂/T₁)), and isothermal expansion of an ideal gas increases entropy (ΔS = nR ln(V₂/V₁)).
  • Common confusion: The same state change can have different paths—reversible vs irreversible—but the entropy change depends only on initial and final states, not the path; however, the relation to heat transfer differs (equality for reversible, inequality for irreversible).
  • Internal heat flow: When heat flows spontaneously within an isolated system from hot to cold regions, entropy increases until thermal equilibrium (uniform temperature) is reached.

📐 The unified mathematical statement

📐 Combining equality and inequality

The excerpt derives a general relation that covers both reversible and irreversible changes:

dS ≥ ∂q/T_b (closed system)

  • The inequality (dS > ∂q/T_b) applies to irreversible changes.
  • The equality (dS = ∂q/T_b) applies to reversible changes.
  • The notation "irrev/rev" indicates which case applies.

🔗 Integrated form

For finite processes:

ΔS ≥ ∫(∂q/T_b) (closed system)

  • This integrates the infinitesimal relation over many cycles.
  • The inequality holds for irreversible processes; the equality holds for reversible processes.

🌡️ Temperature uniformity in reversible processes

  • During a reversible process, states are equilibrium states and temperature is usually uniform throughout the system.
  • Exception: systems with internal adiabatic partitions can have phases at different temperatures in equilibrium.
  • When reversible and temperature is uniform, the relation can be written as dS = ∂q_rev/T (using the system's temperature T instead of boundary temperature T_b).

🔥 Reversible heating and cooling

🔥 Entropy change during heating

The excerpt defines heat capacity C as ∂q/dT (from earlier in the text).

For reversible heating or cooling of a homogeneous phase:

  • ∂q equals T dS
  • The entropy change is:

ΔS = ∫(from T₁ to T₂) (C/T) dT

  • Replace C with C_V (constant volume) or C_p (constant pressure) as appropriate.

📊 Constant heat capacity case

If heat capacity is constant over the temperature range:

ΔS = C ln(T₂/T₁)

  • Heating increases entropy (T₂ > T₁ → positive ΔS).
  • Cooling decreases entropy (T₂ < T₁ → negative ΔS).

Example: An object heated from temperature T₁ to T₂ with constant heat capacity C has entropy increase C ln(T₂/T₁).

🎈 Reversible expansion of an ideal gas

🎈 Adiabatic expansion

When the volume of an ideal gas (or any fluid) is changed reversibly and adiabatically:

  • There is no entropy change (ΔS = 0).
  • No heat is transferred.

🌡️ Isothermal expansion

When the volume of an ideal gas is changed reversibly and isothermally:

  • Expansion work is w = nRT ln(V₂/V₁) (from earlier in the text).
  • Internal energy of an ideal gas is constant at constant temperature.
  • Heat must equal the negative of work: q = -nRT ln(V₂/V₁).
  • The entropy change is:

ΔS = nR ln(V₂/V₁) (reversible isothermal volume change of an ideal gas)

  • Isothermal expansion increases entropy (V₂ > V₁ → positive ΔS).
  • Isothermal compression decreases entropy (V₂ < V₁ → negative ΔS).

🔄 Path independence

  • Since entropy is a state function, the change depends only on initial and final states.
  • The formula ΔS = nR ln(V₂/V₁) is valid for any process with T₂ = T₁, not just reversible isothermal paths.
  • The intermediate states need not be equilibrium states at the same temperature.

Don't confuse: The entropy change is path-independent, but the relation between ΔS and heat transfer depends on whether the actual path is reversible (equality) or irreversible (inequality).

🔒 Spontaneous changes in isolated systems

🔒 What is an isolated system

An isolated system exchanges no matter or energy with its surroundings.

  • Any change that actually occurs is spontaneous: it arises solely from conditions within the system, uninfluenced by the surroundings.
  • The process occurs by itself, of its own accord.

⚡ Characteristics of spontaneous changes

  • The initial and intermediate states must be nonequilibrium states (by definition, an equilibrium state would not change over time).
  • Unless the change is purely mechanical, it is irreversible.

📈 Entropy increase during spontaneous processes

According to the second law, for an irreversible, adiabatic change:

dS > 0 (irreversible change, isolated system)

Key conclusion from the excerpt:

The entropy of an isolated system continuously increases during a spontaneous, irreversible process until it reaches a maximum value at equilibrium.

🌌 Clausius's statement

If we treat the universe as an isolated system:

  • As spontaneous changes occur, the universe's entropy continuously increases.
  • Clausius summarized: "The energy of the universe is constant; the entropy of the universe strives toward a maximum."

🌡️ Internal heat flow in an isolated system

🌡️ Setup and initial state

Consider a solid body with initially nonuniform temperature:

  • The initial state is a nonequilibrium state lacking internal thermal equilibrium.
  • Surrounded by thermal insulation with negligible volume changes → isolated system.
  • Spontaneous, irreversible internal redistribution of thermal energy occurs.
  • Final state: equilibrium with uniform temperature.

🧩 Phase model for analysis

To specify internal temperatures at any instant:

  • Treat the system as an assembly of phases, each with uniform temperature that varies with time.
  • For regions with continuous temperature gradients, approximate with many small phases (parcels), each with temperature infinitesimally different from neighbors.
  • Label phases with Greek letters (φ, ψ, etc.).

🔄 Heat transfer between phases

Let ∂q_φψ represent infinitesimal heat transferred to phase φ from phase ψ:

  • If T_φ < T_ψ (phases in thermal contact), then ∂q_φψ is positive (heat flows to cooler phase).
  • If T_φ > T_ψ (phases in thermal contact), then ∂q_φψ is negative.
  • Otherwise, ∂q_φψ is zero.

📊 Entropy change calculation

The net heat to phase φ during an infinitesimal time interval:

  • ∂q_φ = sum over all ψ≠φ of ∂q_φψ
  • Entropy change of phase φ: dS_φ = ∂q_φ/T_φ (as if heat transferred reversibly from a reservoir at T_φ)

Total entropy change (summing over all phases):

  • After algebraic manipulation using energy conservation (∂q_ψφ = -∂q_φψ):

dS = sum over all pairs φ,ψ of [(1/T_φ - 1/T_ψ) ∂q_φψ]

✅ Why entropy increases

For any nonzero heat transfer term:

  • If T_φ < T_ψ: both ∂q_φψ and (1/T_φ - 1/T_ψ) are positive → term is positive.
  • If T_φ > T_ψ: both ∂q_φψ and (1/T_φ - 1/T_ψ) are negative → term is positive.
  • Each term is either zero or positive.

Conclusion: As long as phases of different temperature are present, dS is positive. Entropy continuously increases until thermal equilibrium (single uniform temperature) is reached.

This agrees with the general principle dS > 0 for irreversible changes in isolated systems.

🎯 Free expansion of a gas

🎯 The process

Setup (referencing an earlier figure in the text):

  • System: the gas.
  • Vessel walls are rigid and adiabatic → system is isolated.
  • Stopcock between two vessels is opened.
  • Gas expands irreversibly into vacuum without heat or work, at constant internal energy.

🔄 Finding the entropy change

To calculate ΔS, use a reversible path between the same initial and final states:

  1. Confine gas at initial volume and temperature in a cylinder-and-piston device.
  2. Expand the gas adiabatically with negative work (using the piston).
  3. Add positive heat to return the internal energy reversibly to its initial value.

Because the reversible path has positive heat, the entropy change is positive.

💡 Key insight

This is an example of an irreversible process in an isolated system where:

  • The actual process has no heat or work.
  • A reversible path between the same initial and final states has both heat and work.

Don't confuse: The entropy change is the same for both paths (state function), but the heat and work differ.

⚙️ Adiabatic process with work

⚙️ Work comparison

General principle (from earlier in the text):

  • An adiabatic process with a given initial equilibrium state and a given change of a work coordinate has the least positive or most negative work in the reversible limit.

🔄 Two-step reversible path

Consider an irreversible adiabatic process with work w_irr.

The same change of state can be accomplished reversibly by:

  1. A reversible adiabatic change of the work coordinate with work w_rev.
  2. Reversible transfer of heat q_rev with no further change (in the work coordinate).

This construction allows calculation of entropy change even when the actual process is irreversible and adiabatic.

26

Summary of the Second Law of Thermodynamics

4.7 Summary

🧭 Overview

🧠 One-sentence thesis

The second law establishes that entropy increases during spontaneous processes in isolated systems, and distinguishes reversible processes (where entropy change equals heat divided by boundary temperature) from irreversible processes (where entropy change exceeds this ratio).

📌 Key points (3–5)

  • Core inequality: For any process in a closed system, dS ≥ δq/T_b, with equality only for reversible processes.
  • Process classification: Every conceivable process is either spontaneous (proceeds naturally at finite rate), reversible (continuous equilibrium states), or impossible.
  • Common confusion: "Spontaneous" and "irreversible" are equivalent terms except for purely mechanical processes (which are idealized and ignored in practice).
  • Entropy as a state function: Because S is a state function, ΔS between two states is the same regardless of path, but the relationship to ∫(δq/T_b) depends on whether the path is reversible or irreversible.
  • Isolated system behavior: In isolated systems, entropy continuously increases during spontaneous processes until thermal equilibrium is reached.

🔤 Key terminology and process types

🔤 Fundamental definitions

Reversible process: A process that proceeds by a continuous sequence of equilibrium states.

Spontaneous process: One that proceeds naturally at a finite rate.

Irreversible process: A spontaneous process whose reverse is impossible.

Purely mechanical process: An idealized process without temperature gradients, and without friction or other dissipative effects, that is spontaneous in either direction.

🔀 Process classification scheme

Process typeCharacteristicsRelationship to other terms
SpontaneousProceeds naturally at finite rateEquivalent to "irreversible" (except for purely mechanical cases)
ReversibleContinuous equilibrium statesNot spontaneous; can go either direction
ImpossibleCannot occur naturallyViolates second law
Purely mechanicalNo temperature gradients, no frictionIdealized; ignored in remaining chapters

Don't confuse: The excerpt explicitly states that purely mechanical processes are a special exception where "spontaneous" does not mean "irreversible"—but these idealized processes are ignored in practical thermodynamics.

📐 Mathematical statement of the second law

📐 The fundamental inequality

For a closed system: dS ≥ δq/T_b

Where:

  • dS = infinitesimal change in entropy (a state function)
  • δq = heat transferred at the boundary
  • T_b = temperature at the boundary where heat is transferred

🔍 Two cases

Reversible process:

  • dS = δq/T_b (equality holds)
  • The entropy change exactly equals the integral of heat divided by boundary temperature

Irreversible (spontaneous) process:

  • dS > δq/T_b (inequality is strict)
  • The entropy change is greater than the integral of heat divided by boundary temperature

🧮 Finite changes between states

For a closed system changing from equilibrium state 1 to equilibrium state 2:

  • Reversible path: ΔS = ∫(δq/T_b)
  • Irreversible path: ΔS > ∫(δq/T_b)

Key insight: Because S is a state function, ΔS between two states is the same regardless of which path is taken. However, the value of ∫(δq/T_b) depends on the path—it equals ΔS only for reversible paths.

🌡️ Thermal equilibration in isolated systems

🌡️ Entropy increase during equilibration

The excerpt derives that during spontaneous thermal equilibration in an isolated system:

  • Starting with any initial distribution of internal temperatures
  • Entropy continuously increases
  • Process continues until a single uniform temperature is reached throughout
  • Each term in the mathematical sum is either zero or positive, making dS positive

Example: An isolated system with hot and cold regions will spontaneously equilibrate to uniform temperature, with entropy increasing throughout the process.

⚖️ Final equilibrium state

At thermal equilibrium:

  • Single uniform temperature throughout the system
  • Entropy has reached its maximum value for the isolated system
  • No further spontaneous changes occur

🎈 Illustrative examples from the excerpt

🎈 Free expansion of a gas

Setup: Gas in a vessel with rigid, adiabatic walls (isolated system); stopcock opened to allow expansion into vacuum.

Irreversible process:

  • Gas expands into vacuum
  • No heat transfer (q = 0)
  • No work (w = 0)
  • Constant internal energy

Reversible path for same state change:

  1. Confine gas in cylinder-and-piston at initial volume and temperature
  2. Expand adiabatically with negative work
  3. Add positive heat to return internal energy to initial value

Conclusion: Because the reversible path has positive heat, the entropy change is positive. This demonstrates an irreversible process in an isolated system where the reversible path between the same initial and final states requires both heat and work.

⚙️ Adiabatic process with work

General principle (from page 98): An adiabatic process with given initial state and given work coordinate change has the least positive or most negative work in the reversible limit.

Analysis:

  • Irreversible adiabatic process: work = w_irr
  • Reversible path for same state change:
    1. Reversible adiabatic change of work coordinate: work = w_rev
    2. Reversible heat transfer q_rev with no further work change

Key relationship:

  • w_rev is algebraically less than w_irr
  • To make ΔU the same in both paths, q_rev must be positive
  • Positive heat increases entropy along the reversible path
  • Therefore, the irreversible adiabatic process has positive entropy change

Agreement: This conclusion agrees with the second-law inequality dS ≥ δq/T_b (Eq. 4.6.1).

🔑 Critical distinctions

🔑 Reversible vs irreversible paths

Don't confuse the path with the state change:

  • ΔS (entropy change) depends only on initial and final states (state function property)
  • ∫(δq/T_b) depends on the specific path taken
  • For reversible paths: ΔS = ∫(δq/T_b)
  • For irreversible paths: ΔS > ∫(δq/T_b)

🔑 Heat, work, and entropy in different processes

Process typeHeat (q)Work (w)Entropy change (ΔS)
Free expansion (isolated)00Positive
Reversible path for same changePositiveNegativeSame positive value
Irreversible adiabatic0w_irrPositive
Reversible path for same changePositive (q_rev)w_rev (less than w_irr)Same positive value

Key insight: The same state change (same ΔS) can be achieved by different combinations of heat and work, depending on whether the path is reversible or irreversible.

27

The Statistical Interpretation of Entropy

4.8 The Statistical Interpretation of Entropy

🧭 Overview

🧠 One-sentence thesis

Statistical mechanics interprets entropy as a measure of the number of accessible microstates, meaning entropy increases when energy spreads into more possible microscopic configurations rather than when "disorder" increases.

📌 Key points (3–5)

  • Why "disorder" is misleading: The disorder metaphor fails to explain reversible heating, isothermal expansion, or spontaneous crystallization in isolated systems.
  • What entropy really measures: The statistical entropy equals k ln W, where W is the number of accessible microstates (quantum states) compatible with the system's constraints.
  • How microstates work: A macroscopic equilibrium state is a time average over many microstates; the system continuously jumps between accessible microstates of equal (or nearly equal) energy.
  • Common confusion: Entropy increase does not mean "more disorder"—it means energy is spreading or dispersing into more accessible microstates.
  • Connection to macroscopic entropy: Changes in statistical entropy match changes in second-law entropy because both are derived from the same thermodynamic relation dS = δq/T.

🚫 Why the disorder metaphor fails

🚫 What the disorder idea claims

  • A common description says entropy measures "disorder" on a microscopic scale.
  • According to this view, entropy increases whenever a closed system becomes more disordered.

❌ Where the disorder metaphor breaks down

The excerpt identifies three failures:

ProcessWhat happensWhy disorder fails
Reversible heating at constant volume or pressureEntropy increasesDisorder metaphor does not explain why adding heat reversibly increases entropy
Reversible isothermal expansion of ideal gasEntropy increasesDisorder metaphor does not explain why volume increase at constant temperature increases entropy
Freezing of supercooled liquid or crystallization from supersaturated solutionEntropy decreases, yet process is spontaneous in isolated systemApparent decrease of disorder contradicts the idea that spontaneous processes always increase disorder
  • Key takeaway: The excerpt states explicitly, "we should not interpret entropy as a measure of disorder."
  • Don't confuse: spontaneous processes in isolated systems must increase total entropy, but local entropy can decrease (e.g., crystallization) as long as the system is not perfectly isolated or the total entropy accounting is correct.

🔬 The statistical mechanics foundation

🔬 What statistical mechanics provides

Statistical mechanics: the discipline that derives a precise expression for entropy based on the behavior of macroscopic amounts of microscopic particles.

  • It offers a rigorous microscopic interpretation of entropy.
  • The excerpt contrasts this with the vague and misleading disorder metaphor.

⚛️ Microstates and macrostates

  • Macroscopic equilibrium state: what classical thermodynamics describes with state functions like temperature, pressure, volume.
  • Microstate (stationary quantum state): a specific quantum configuration with definite energy that the system occupies at a given instant.
  • Over time, the system in an equilibrium state continually jumps from one microstate to another.
  • Macroscopic state functions are time averages of these microstates.

🔓 Accessible microstates

Accessible microstate: a microstate whose wave function is compatible with the system's volume and any other conditions and constraints imposed on the system.

  • Not all microstates are accessible—only those that satisfy the system's constraints (volume, energy, etc.).
  • Example: If a gas is confined to a certain volume, only microstates with wave functions fitting that volume are accessible.

⚖️ The fundamental assumption

  • Equal probability: Accessible microstates of equal energy are equally probable.
  • The system spends an equal fraction of its time in each such microstate.
  • This assumption is the foundation of statistical mechanics.

📐 The statistical entropy formula

📐 The equation

The statistical entropy is given by:

S_stat = k ln W + C

where:

  • k is the Boltzmann constant (k = R / N_A, where R is the gas constant and N_A is Avogadro's number)
  • W is the number of accessible microstates
  • C is a constant

🔋 Two scenarios for accessible microstates

🔋 Perfectly isolated system (microcanonical ensemble)

  • The system has constant internal energy U.
  • Accessible microstates are those compatible with constraints and whose energies all equal U.
  • All accessible microstates have exactly the same energy.

🌡️ System in thermal equilibrium with reservoir (canonical ensemble)

  • The system is at constant temperature T.
  • Internal energy fluctuates over time with extremely small deviations from average value U.
  • Accessible microstates have energies close to this average value.
  • The excerpt notes this is "more realistic."

🔗 Connection to macroscopic entropy

  • Changes match: A change ΔS_stat equals the change ΔS of the macroscopic second-law entropy.
  • Why they match: The derivation of the statistical formula is based on the macroscopic relation dS_stat = δq/T = (dU − δw)/T, with dU and δw given by statistical theory.
  • The integration constant: If C is set to zero, S_stat becomes the third-law entropy S (to be described in Chapter 6).

🌊 Energy spreading and dispersal

🌊 What entropy increase really means

The excerpt states that Equation 4.8.1 shows:

A reversible process in which entropy increases is accompanied by an increase in the number of accessible microstates of equal, or nearly equal, internal energies.

  • More accessible microstates → higher entropy.
  • Fewer accessible microstates → lower entropy.

📖 Alternative descriptions

The excerpt mentions several ways this interpretation has been described:

DescriptionWhat it emphasizes
Spreading and sharing of energyEnergy becomes distributed among more microstates
Dispersal of energyEnergy spreads out into more configurations
"Spreading function"Entropy as S suggesting "spreading"
  • Example: In isothermal expansion of an ideal gas, the volume increases, so more spatial configurations (microstates) become accessible → W increases → entropy increases.
  • Don't confuse: "spreading" refers to the number of accessible microstates, not to physical scattering or spatial disorder in the everyday sense.

🔤 The symbol S

  • The excerpt notes that the symbol S for entropy was originally an arbitrary choice by Clausius.
  • The "spreading function" interpretation is a modern pedagogical proposal, not the historical origin.
28

Total Differential of a Dependent Variable

5.1 Total Differential of a Dependent Variable

🧭 Overview

🧠 One-sentence thesis

The excerpt does not contain substantive content on the total differential of a dependent variable; instead, it presents material on reversible and irreversible processes, work calculations, and energy dissipation in thermodynamic systems.

📌 Key points (3–5)

  • Mismatch between title and content: The title suggests mathematical treatment of total differentials, but the excerpt covers thermodynamic processes and work.
  • Core thermodynamic concepts present: reversible vs. irreversible work, energy dissipation, adiabatic processes, and work coordinates.
  • Work formulas provided: The excerpt includes a table of work expressions for various physical processes (expansion, electrical, mechanical, etc.).
  • Common confusion addressed: The difference between lubricated and dry friction in irreversible processes, and how work magnitude differs between expansion and compression.
  • State function emphasis: Work coefficients and coordinates for reversible work are state functions, but dissipative work coordinates are not.

🔄 Reversible vs. Irreversible Processes

🔄 Core distinction in work recovery

Energy transferred across the boundary by work in a reversible process is fully recovered as work of the opposite sign in the reverse reversible process.

  • In a reversible process, you can "undo" the work completely—if you compress a gas reversibly, you can expand it reversibly and recover all the work.
  • The first law implies that heat is also fully recovered in the reverse process.
  • Example: A reversible adiabatic expansion followed by a reversible adiabatic compression returns the system to its original state with no net work or heat.

⚠️ Irreversible work and energy dissipation

  • When work occurs irreversibly at a finite rate, there is partial or complete dissipation of energy.
  • The dissipation results in a change that could also be accomplished with positive heat, such as an increase of thermal energy within the system.
  • Don't confuse: Dissipation doesn't mean energy is "lost"—it means energy is converted to a form (like thermal energy) that cannot be fully recovered as work.

🎯 Adiabatic work limits

For adiabatic processes with a reversible limit:

  • The work for a given initial equilibrium state and a given change in the work coordinate is least positive or most negative in the reversible limit.
  • This means:
    • For expansion: reversible work is more negative (you extract more work) than irreversible work.
    • For compression: reversible work is less positive (you do less work) than irreversible work.
  • The excerpt confirms this with indicator diagrams showing that irreversible expansion requires less work magnitude than reversible, while irreversible compression requires more work magnitude.

🔧 Friction Effects in Thermodynamic Processes

🛢️ Lubricated friction behavior

  • With lubricated friction (e.g., a rod sliding in a lubricated bushing), the frictional force approaches zero in the limit of zero piston velocity.
  • As the rate of volume change approaches zero, the process approaches the reversible limit.
  • The work of expansion and compression approach the same magnitude (but opposite signs) in the reversible limit.
  • Example: Figure 3.21 shows that as the average rate of volume change decreases, both expansion and compression work approach the reversible limit values (shown as open circles).

🪨 Dry friction behavior

  • With dry friction (unlubricated surfaces in direct contact), the frictional force does not approach zero even at zero piston velocity.
  • Dry friction is due to microscopic surface roughness and is typically independent of contact area and sliding rate.
  • In the limit of infinite slowness, the expansion work is smaller in magnitude than the compression work.
  • The internal dry friction prevents the expansion process from being reversed as a compression process, regardless of piston velocity—these processes are therefore irreversible.
  • Don't confuse: Unlike lubricated friction, dry friction does not allow a reversible limit even at infinitely slow rates.

🔥 Energy dissipation details

The thermal energy released by frictional work increases the gas pressure, making:

  • For expansion: the difference between irreversible and reversible work (w_irr minus w_rev) less than the frictional work itself.
  • For compression: the difference between irreversible and reversible work greater than the frictional work itself.
  • The excerpt notes there is no general method to evaluate energy dissipated by internal friction, especially for processes with both work and heat.

📐 Work Coordinates and State Functions

📐 General work formula

Infinitesimal quantities of work during a process are calculated from an expression of the form: δw = sum over i of (Y_i times dX_i), where X_i is the work coordinate of kind of work i and Y_i is the conjugate work coefficient.

  • X_i represents the extensive quantity that changes (e.g., volume, length, angle).
  • Y_i represents the intensive quantity that drives the work (e.g., pressure, force, torque).
  • Example: For expansion work, X is volume V, Y is boundary pressure p_b, so δw = p_b dV.

🎯 State function properties

  • The work coefficients and work coordinates of reversible work are state functions.
  • This means their values depend only on the current equilibrium state, not on how the system reached that state.
  • However, dissipative work coordinates are not state functions.

🔢 Independent variables for equilibrium states

The number of independent variables needed to describe equilibrium states of a closed system is:

  • One greater than the number of independent work coordinates for reversible work.
  • You could choose the independent variables to be each of the work coordinates plus either the temperature or the internal energy.
  • Example: For a simple gas with only expansion work (one work coordinate, volume), you need two variables: volume and temperature (or volume and internal energy).
  • Don't confuse: Nonequilibrium states require many more independent variables than equilibrium states.

📊 Types of Work in Thermodynamic Systems

📊 Dissipative work

Dissipative work is positive irreversible work with complete energy dissipation. The work coordinate for this type of work is not a state function.

  • Examples from the excerpt: stirring work and electrical heating work.
  • All the work input is converted to thermal energy; none can be recovered as work.
  • The work coordinate is not a state function because the amount of stirring or electrical heating done depends on the path, not just the initial and final states.

📊 Work with partial energy dissipation

  • Some processes have partial energy dissipation—not all work is converted to thermal energy.
  • Examples: adiabatic expansion/compression with lubricated friction, galvanic cell electrical work.
  • The dependence of work on the rate of change is shown graphically in Figures 3.12(a), 3.18, and 3.21.

📋 Common work formulas

The excerpt provides a table of work expressions for various processes:

Kind of workFormulaKey variables
Linear mechanicalδw = F_sur,x dxF_sur,x = x-component of force by surroundings; dx = displacement
Shaft workδw = τ_b dθτ_b = internal torque at boundary; θ = angle of rotation
Expansion workδw = p_b dVp_b = average pressure at moving boundary; V = volume
Surface workδw = γ dA_sγ = surface tension; A_s = surface area
Stretching/compressionδw = F dlF = stress (positive for tension, negative for compression); l = length
Gravitational workδw = mg dhm = mass; g = acceleration of free fall; h = height
Electrical workδw = Φ dQ_sysΦ = electric potential difference; Q_sys = charge entering system
Electric polarizationδw = E · dpE = electric field strength; p = electric dipole moment
Magnetizationδw = B · dmB = magnetic flux density; m = magnetic dipole moment

Note: The dot (·) represents a dot product for vector quantities.

29

Total Differential of the Internal Energy

5.2 Total Differential of the Internal Energy

🧭 Overview

🧠 One-sentence thesis

The excerpt does not contain substantive content about the total differential of internal energy; instead, it presents thermodynamics problems and introduces the second law of thermodynamics, focusing on process types and entropy.

📌 Key points (3–5)

  • The excerpt contains problem sets from Chapter 3 (First Law) involving calculations of work, heat, and internal energy changes.
  • Chapter 4 introduces the second law of thermodynamics, which addresses spontaneity and entropy rather than internal energy differentials.
  • Three process types are defined: spontaneous (actually occurs), reversible (idealized equilibrium sequence), and impossible (cannot occur).
  • The mathematical statement of the second law relates entropy change to heat transfer: dS = δq/T_b for reversible processes, dS > δq/T_b for irreversible processes.
  • Common confusion: The excerpt does not address the topic indicated by the title "5.2 Total Differential of the Internal Energy."

📋 Content mismatch

📋 What the excerpt contains

The provided text includes:

  • Chapter 3 problems: Numerical exercises about gas processes, piston-cylinder systems, adiabatic compression, Joule's paddle wheel experiment, and mechanical equivalent of heat calculations.
  • Chapter 4 introduction: Theoretical discussion of the second law, process classifications, and the mathematical definition of entropy.

❌ What is missing

  • No section numbered 5.2.
  • No discussion of the total differential of internal energy (typically written as dU = TdS - pdV or similar expressions).
  • No derivation or explanation of how internal energy depends on its natural variables.

🔄 Process types (from Chapter 4)

🔄 Three fundamental categories

The excerpt defines three mutually exclusive process types:

Process typeDefinitionCan it occur?
SpontaneousA real process that can take place in finite timeYes, actually happens
ReversibleIdealized process through continuous equilibrium statesNo, but can be approached
ImpossibleCannot occur under existing conditions, even as a limitNo, never

⚡ Irreversible processes

An irreversible process is a spontaneous process whose reverse is an impossible process.

  • Most spontaneous processes relevant to chemistry are irreversible.
  • Example context from excerpt: phase changes, solute transfers, chemical reactions under given conditions.

🔧 Purely mechanical processes

A purely mechanical process is a spontaneous process whose reverse is also spontaneous.

  • Special category of little interest to chemists.
  • Can be "practically" achieved when friction and temperature gradients are negligible.

🌡️ The second law introduction

🌡️ Mathematical statement of entropy

The excerpt presents the core mathematical formulation:

dS = δq/T_b for a reversible change of a closed system;
dS > δq/T_b for an irreversible change of a closed system;

Where:

  • S is entropy, an extensive state function
  • δq is an infinitesimal quantity of energy transferred by heat
  • T_b is the thermodynamic temperature at the boundary portion where heat transfer occurs

🎯 Three components of the statement

  1. Assertion: Entropy S exists as an extensive state function.
  2. Equation: Provides a method for calculating entropy change in a closed system during a reversible process.
  3. (Implied third part): The inequality for irreversible processes establishes the direction of spontaneous change.

🔍 Why the second law matters

  • The first law (energy balance) cannot predict which processes occur spontaneously.
  • Physics explains some spontaneous changes (unbalanced forces, temperature gradients).
  • The second law provides a general criterion for spontaneity in closed systems, answering questions about phase changes, solute transfers, and chemical reactions.

⚠️ Note on content relevance

The excerpt does not contain material matching the title "5.2 Total Differential of the Internal Energy." The problems and theoretical content present are from Chapters 3 and 4 of a thermodynamics textbook, covering the first and second laws respectively, but not the specific mathematical treatment of internal energy's total differential that would typically appear in a section with that title.

30

Enthalpy, Helmholtz Energy, and Gibbs Energy

5.3 Enthalpy, Helmholtz Energy, and Gibbs Energy

🧭 Overview

🧠 One-sentence thesis

The second law of thermodynamics establishes entropy as a state function and provides a general criterion for determining whether processes will occur spontaneously under given conditions.

📌 Key points (3–5)

  • What the second law addresses: the first law accounts for energy changes but cannot predict spontaneity; the second law fills this gap by introducing entropy and spontaneity criteria.
  • Three process types: any conceivable process is either spontaneous (actually occurs), reversible (idealized equilibrium sequence), or impossible (cannot occur even in principle).
  • Two equivalent physical statements: the Clausius statement (heat cannot flow from cold to hot without other changes) and the Kelvin–Planck statement (a heat engine cannot convert heat entirely to work in a cycle) both assert certain processes are impossible despite conserving energy.
  • Common confusion: impossible processes do not violate the first law (energy is conserved), but they violate the second law—time-reversal symmetry at the molecular level does not guarantee macroscopic reversibility.
  • Mathematical formulation: entropy S is an extensive state function; for reversible changes dS = δq/Tₑ, for irreversible changes dS > δq/Tₑ, providing the spontaneity criterion.

🔄 Process classification and spontaneity

🔄 Three types of processes

The excerpt categorizes all conceivable processes into three mutually exclusive types:

Process typeDefinitionCan it occur?
SpontaneousA real process that can actually take place in a finite time periodYes, in practice
ReversibleAn imaginary process passing through continuous equilibrium states; can be approached by infinitely slow spontaneous processesOnly as an idealized limit
ImpossibleA change that cannot occur under existing conditions, even in a limiting sense (also called unnatural or disallowed)No, not even in principle
  • The excerpt emphasizes that reversible processes are "imaginary, idealized" but useful for theoretical analysis.
  • Example: A gas expanding infinitely slowly through equilibrium states is reversible; rapid expansion is spontaneous and irreversible.

⚙️ Irreversible vs purely mechanical processes

An irreversible process is a spontaneous process whose reverse is an impossible process.

  • Most spontaneous processes relevant to chemistry are irreversible—once they occur, the reverse cannot happen spontaneously.
  • Purely mechanical processes: a special category where both the process and its reverse are spontaneous (e.g., frictionless motion); these are "of little interest to chemists."
  • Don't confuse: "practically reversible" (a spontaneous process carried out very slowly) is not the same as truly impossible—the excerpt reserves "impossible" for processes that cannot be approached no matter how carefully conducted.

🚫 Physical statements of the second law

🚫 Why certain processes are impossible

The excerpt notes that impossible processes do not violate the first law (energy is conserved) but are nevertheless never observed. Two equivalent physical principles describe this:

❄️ The Clausius statement

It is impossible to construct a device whose only effect, when it operates in a cycle, is heat transfer from a body to the device and the transfer by heat of an equal quantity of energy from the device to a warmer body.

  • Plain language: Heat cannot spontaneously flow from cold to hot without some other change occurring.
  • Why it matters: Figure 4.1(a) shows an isolated system where heat flows from cool to warm body—this is impossible even though energy is conserved.
  • With a device (Figure 4.1(b)): Even if you insert a device operating in a cycle to transfer heat from cold to hot, it cannot do so without some other effect—the device cannot return to its initial state while only moving heat "uphill."
  • Example: You cannot build a refrigerator that cools something without consuming work or producing another change.

🔥 The Kelvin–Planck statement

It is impossible to construct a heat engine whose only effect, when it operates in a cycle, is heat transfer from a heat reservoir to the engine and the performance of an equal quantity of work on the surroundings.

  • Plain language: You cannot build an engine that converts heat entirely into work in a cycle with no other effect.
  • Why it matters: Figure 4.2(b) shows a heat engine that would extract heat from a reservoir (e.g., the ocean) and convert it all to work (raising a weight), then return to its initial state—this is impossible.
  • The excerpt notes this would be "very desirable" (a ship powered by ocean heat with no fuel) but cannot be done.
  • Don't confuse: The engine can convert some heat to work, just not all of it without other changes (this is why real engines need both a hot and cold reservoir).

🔗 Equivalence of the two statements

The excerpt states that the Clausius and Kelvin–Planck statements are equivalent: "if one is true, so is the other." Section 4.3 will prove this equivalence using a hypothetical Carnot engine.

📐 Mathematical formulation of the second law

📐 Three components of the mathematical statement

The box on page 106 gives the formal second law in three parts:

  1. Entropy exists: S is an extensive state function called entropy.
  2. Reversible change equation: For a reversible change of a closed system, dS = δq/Tₑ, where δq is an infinitesimal quantity of heat transferred at a boundary portion with thermodynamic temperature Tₑ.
  3. Spontaneity criterion: For an irreversible change of a closed system, dS > δq/Tₑ.
  • The temperature Tₑ is "thermodynamic temperature" (defined in Sec. 4.3.4).
  • The excerpt acknowledges these statements are "somewhat abstract" and asks: "What fundamental principle, based on experimental observation, may we take as the starting point to obtain them?"
  • Answer: The Clausius and Kelvin–Planck statements (impossible-process principles) serve as the experimental foundation.

🧮 Special case for uniform temperature

The footnote clarifies: "During a reversible process, the temperature usually has the same value T throughout the system, in which case we can simply write dS = δq/T."

  • The more general form dS = δq/Tₑ allows for phases at different temperatures separated by internal adiabatic partitions in an equilibrium state.

🔍 Why entropy is not obvious from energy conservation

  • The excerpt emphasizes that impossible processes (Figures 4.1 and 4.2) do not violate the first law—energy is conserved in both.
  • Example: In Figure 4.2(a), a weight rises while water cools; potential energy increases exactly as thermal energy decreases—first law is satisfied, yet the process is impossible.
  • The second law introduces a new criterion (entropy change) beyond energy accounting.

🔬 Microscopic perspective and time reversal

🔬 Why macroscopic irreversibility despite microscopic reversibility

The excerpt addresses a fundamental puzzle:

Newton's laws of motion are invariant to time reversal. Suppose we could measure the position and velocity of each molecule of a macroscopic system in the final state of an irreversible process. Then, if we could somehow arrange at one instant to place each molecule in the same position with its velocity reversed, and if the molecules behaved classically, they would retrace their trajectories in reverse and we would observe the reverse "impossible" process.

  • Plain language: At the molecular level, the laws of physics work the same forward and backward in time.
  • The puzzle: If we could reverse every molecule's velocity, an "impossible" macroscopic process (like heat flowing from cold to hot) would occur.
  • Implication: Macroscopic irreversibility arises not from fundamental physics but from the practical impossibility of controlling vast numbers of particles—this hints at the statistical nature of entropy (discussed in Sec. 4.8).
  • Don't confuse: Microscopic time-reversal symmetry does not mean macroscopic processes are reversible; the second law is a statistical statement about large ensembles.

🛠️ Carnot engines and derivation strategy

🛠️ Why Carnot cycles are used

The excerpt acknowledges that "Carnot engines and Carnot cycles are admittedly outside the normal experience of chemists" and quotes Lewis and Randall's 1923 complaint about "'cyclical processes' limping about eccentric and not quite completed cycles."

  • Why necessary: "There seems, however, to be no way to carry out a rigorous general derivation without invoking thermodynamic cycles."
  • The excerpt offers an escape: "You may avoid the details by skipping Secs. 4.3–4.5."

🔧 What a Carnot engine is

A Carnot engine is a particular kind of heat engine, one that performs Carnot cycles with a working substance.

  • Heat engine definition: A closed system that converts heat to work and operates in a cycle.
  • Carnot cycle: Four reversible steps, alternating isothermal and adiabatic.

🔧 The four steps of a Carnot cycle

The excerpt describes the cycle (illustrated in Figures 4.3 and 4.4 for ideal gas and water):

  1. Path A → B (isothermal heat absorption): Heat qₕ is transferred reversibly and isothermally from a hot reservoir at temperature Tₕ to the system; qₕ is positive (energy into system).
  2. Path B → C (adiabatic expansion): The system undergoes reversible adiabatic change, does work on surroundings, and temperature drops to Tₑ.
  3. (Paths C → D and D → A not fully described in excerpt but implied by cycle structure): The cycle returns to initial state.
  • Example (Figure 4.3): For an ideal gas with Tₕ = 400 K, Tₑ = 300 K, the cycle A → B → C → D → A has net work w = 1.0 J.
  • The excerpt notes that Section 4.3 will use Carnot engines to prove equivalence of the two second-law statements and to define thermodynamic temperature via engine efficiency.

📚 Derivation roadmap

📚 Structure of the remaining chapter

The excerpt outlines the logical flow:

  • Sec. 4.3: Introduce Carnot engine; prove Clausius and Kelvin–Planck statements are equivalent; derive Carnot efficiency to define thermodynamic temperature.
  • Sec. 4.4: Combine Carnot cycles and Kelvin–Planck statement to derive existence and properties of entropy (the state function S).
  • Sec. 4.5: Use irreversible processes to complete the derivation of the mathematical statements (dS = δq/Tₑ for reversible, dS > δq/Tₑ for irreversible).
  • Sec. 4.6: Applications of the second law.
  • Sec. 4.7: Summary.
  • Sec. 4.8: Microscopic, statistical interpretation of entropy.

This roadmap shows how the abstract mathematical formulation is built rigorously from the experimentally grounded impossible-process statements.

31

Closed Systems

5.4 Closed Systems

🧭 Overview

🧠 One-sentence thesis

Carnot engines—idealized closed systems operating in reversible cycles between two heat reservoirs—provide the logical foundation for proving the equivalence of the second law's physical statements and deriving the concept of entropy.

📌 Key points (3–5)

  • What a Carnot engine is: a closed system that converts heat to work through a four-step reversible cycle (alternating isothermal and adiabatic processes) between two fixed temperatures.
  • Why Carnot cycles matter: they allow rigorous general derivation of the second law's mathematical statement, despite being outside normal chemistry experience.
  • Key energy flow: heat enters from the hot reservoir, part is converted to work, and the remainder is transferred to the cold reservoir (preventing violation of the Kelvin–Planck statement).
  • Reversibility: all four steps are reversible processes, so the cycle can run forward (as an engine) or backward (as a heat pump); neither direction is impossible.
  • Common confusion: Carnot engines are idealized and do not resemble practical steam engines, which use open systems and irreversible steps.

🔄 The Carnot cycle structure

🔄 Four reversible steps

A Carnot cycle consists of four alternating processes (the system is the working substance):

StepProcessWhat happensHeat/work
A → BIsothermalHeat q_h transferred from hot reservoir at T_h to systemq_h positive (into system)
B → CAdiabaticSystem does work on surroundings, temperature drops to T_cNo heat transfer
C → DIsothermalHeat q_c transferred from system to cold reservoir at T_cq_c negative (out of system)
D → AAdiabaticWork done on system, temperature returns to T_h, system returns to initial stateNo heat transfer
  • The cycle is complete: the system returns to its initial state after step D → A.
  • Adjusting the length of path A → B changes the magnitude of net work w (can be made as large or small as desired).

🌡️ Temperature requirements

  • The cycle operates between two fixed temperatures: T_h (hot reservoir) and T_c (cold reservoir).
  • During isothermal steps, the system temperature matches the reservoir temperature for reversible heat transfer.
  • During adiabatic steps, the system temperature changes between T_h and T_c with no heat transfer.

⚙️ Energy flow and work output

⚙️ Net energy balance

In one cycle: a quantity of heat is transferred from the hot reservoir to the system, a portion of this energy is transferred as heat to the cold reservoir, and the remainder of the energy is the negative net work w done on the surroundings.

  • Energy in: q_h from hot reservoir
  • Energy out: |q_c| to cold reservoir + |w| as work to surroundings
  • The heat transfer to the cold reservoir is what keeps the Carnot engine from being an impossible Kelvin–Planck engine (which would convert heat entirely to work with no other effect).

🔁 Carnot heat pump (reverse cycle)

When the cycle runs in reverse:

  • q_h becomes negative (heat transferred from system to hot reservoir)
  • q_c becomes positive (heat transferred from cold reservoir to system)
  • Work must be done on the system to drive the reverse cycle
  • This is called a Carnot heat pump
  • Since each step is reversible, neither the forward engine nor the reverse heat pump is impossible.

🏭 Carnot vs. practical engines

🏭 Why Carnot engines are idealized

  • Reversibility: All paths are reversible processes—this is an idealization not achievable in real engines.
  • Closed system: The Carnot engine keeps the same working substance throughout the cycle.
  • Purpose: The Carnot engine is a hypothetical device used for rigorous derivation, not a practical design.

🚂 Practical steam engines (contrast)

Real working steam engines differ fundamentally:

  • Use an open system (steam enters and exits the cylinder)
  • Undergo irreversible steps in each cycle:
    1. High-pressure steam enters from boiler and pushes piston
    2. Supply valve closes, steam expands until pressure drops to atmospheric
    3. Exhaust valve opens to release steam (to atmosphere or condenser)
    4. Piston returns to initial position (driven by external force or suction from condensation)
  • Do not resemble Carnot engine design at all

Don't confuse: Carnot engines are theoretical tools for deriving thermodynamic principles; practical engines sacrifice reversibility for real-world operation.

🧪 Working substances and examples

🧪 Ideal gas example

The excerpt describes a Carnot cycle using an ideal gas (Figure 4.3):

  • T_h = 400 K, T_c = 300 K
  • Paths A → B and C → D are isothermal
  • Paths B → C and D → A are adiabatic
  • Two different cycles shown with net work w = 1.0 J and w = 0.5 J (by adjusting path A → B length)

💧 Water example

The excerpt describes a Carnot cycle using H₂O (Figure 4.4):

  • T_h = 400 K, T_c = 396 K
  • Initial state A: one mole of liquid water
  • A → B: reversible vaporization of 2.54 mmol at 400 K
  • B → C: adiabatic expansion causing additional 7.68 mmol to vaporize
  • C → D: reversible condensation of 2.50 mmol at 396 K
  • D → A: adiabatic compression returning to initial state
  • Net work w = 1.0 J

These examples show Carnot cycles work with different substances and involve phase changes (not just pressure-volume work).

🎯 Purpose in thermodynamic derivation

🎯 Why use Carnot cycles

The excerpt acknowledges that "Carnot engines and Carnot cycles are admittedly outside the normal experience of chemists" and that using them "may seem arcane."

However:

  • There is "no way to carry out a rigorous general derivation without invoking thermodynamic cycles."
  • Carnot cycles are used to:
    • Prove the equivalence of the Clausius and Kelvin–Planck statements (Section 4.3)
    • Define thermodynamic temperature via engine efficiency
    • Derive the existence and properties of entropy (Section 4.4)
    • Complete the mathematical statement of the second law (Section 4.5)

🔗 Logical tool: reductio ad absurdum

The excerpt introduces the method for proving equivalence of the two second-law statements:

  • Assume one statement (e.g., Clausius) is incorrect
  • Show that if the "impossible" device were possible, it could be combined with a Carnot engine
  • Demonstrate this combination would create another impossible device (violating the other statement)
  • Therefore, if one statement is true, the other must also be true

Example setup: combine a hypothetical "Clausius device" with a Carnot engine, adjusting cycles so equal heat quantities are transferred to/from the cold reservoir.

Why this matters: This logical structure establishes that the Clausius and Kelvin–Planck statements are equivalent—they are two ways of expressing the same fundamental law.

32

Open Systems

5.5 Open Systems

🧭 Overview

🧠 One-sentence thesis

The Carnot engine is an idealized reversible heat engine that demonstrates fundamental thermodynamic principles, while real steam engines operate as open systems with irreversible processes that differ fundamentally in design and operation.

📌 Key points (3–5)

  • Carnot engine idealization: The Carnot engine uses reversible processes and does not resemble any practical steam engine design.
  • Real steam engines as open systems: Practical steam engines contain open systems that undergo irreversible steps involving steam entering, expanding, exhausting, and piston return.
  • Carnot heat pump: Running a Carnot engine in reverse creates a heat pump where heat flow directions are reversed.
  • Common confusion: Don't confuse the idealized Carnot engine (closed, reversible) with practical steam engines (open, irreversible)—they operate on fundamentally different principles despite both being "heat engines."
  • Equivalence of second law statements: The Clausius and Kelvin–Planck statements of the second law are logically equivalent and can be proven using reductio ad absurdum.

🔄 Carnot engine vs practical steam engines

🎯 The idealized Carnot engine

Carnot engine: an idealized heat engine whose paths are reversible processes.

  • The Carnot engine operates through four reversible steps involving heat transfer and adiabatic changes.
  • Path A→B: Reversible isothermal heat transfer from hot reservoir at temperature T_h.
  • Path B→C: Reversible adiabatic expansion.
  • Path C→D: Reversible isothermal heat transfer to cold reservoir at temperature T_c (q_c is negative).
  • Path D→A: Reversible adiabatic compression returning to initial state.
  • The magnitude of work w can be adjusted by changing the length of path A→B.

🚂 Practical steam engines as open systems

Real steam engines differ fundamentally from Carnot engines:

  • The cylinder contains an open system (not closed like Carnot).
  • Each cycle involves irreversible steps:
    1. High-pressure steam enters from boiler, pushes piston
    2. Supply valve closes, steam expands until pressure drops to atmospheric
    3. Exhaust valve opens, steam released to atmosphere or condenser
    4. Piston returns to initial position (driven by external force or suction)

Don't confuse: The Carnot engine is a theoretical model with reversible processes in a closed system; practical steam engines are open systems with irreversible processes and completely different mechanical designs.

Example: A train locomotive steam engine continuously takes in new steam and exhausts it, unlike the Carnot engine which cycles the same working substance reversibly.

🔁 Carnot heat pump (reversed cycle)

🔁 How reversal works

When the Carnot engine cycle is reversed, it becomes a Carnot heat pump:

  • In the heat pump, q_h is negative (heat transferred from system to hot reservoir).
  • q_c is positive (heat transferred to system from cold reservoir).
  • Work must be done on the system (not by the system).
  • Each step remains reversible, so neither the engine nor heat pump is an "impossible device."

📊 Energy transfer comparison

Deviceq_hq_cWorkDirection
Carnot enginePositive (heat in)Negative (heat out)Done by systemHot→System→Cold + Work out
Carnot heat pumpNegative (heat out)Positive (heat in)Done on systemCold→System→Hot (requires Work in)

Example: Figure 4.5(a) shows 4 joules from hot reservoir, 3 joules to cold reservoir, 1 joule of work done on surroundings; reversed in (b), the directions flip.

🔗 Equivalence of second law statements

🧩 The two statements

The excerpt discusses proving that the Clausius and Kelvin–Planck statements are equivalent:

  • Clausius statement: Claims a certain device (spontaneous heat transfer from cold to hot) is impossible.
  • Kelvin–Planck statement: Claims a certain heat engine (converting heat entirely to work with no other change) is impossible.

🔍 Proof by reductio ad absurdum (Clausius → Kelvin–Planck)

The logical tool of reductio ad absurdum (proof by contradiction) establishes equivalence:

Step 1: Assume the Clausius statement is incorrect (the "Clausius device" is possible).

Step 2: Combine the Clausius device with a Carnot engine:

  • Adjust cycles so equal heat quantities transfer from/to the cold reservoir.
  • The combination becomes a single system.

Step 3: After one cycle of each device:

  • Net result: heat transfer into system, equal work done on surroundings, no other net change.
  • This creates a heat engine that the Kelvin–Planck statement says is impossible.

Conclusion: If Kelvin–Planck is correct, the Clausius device cannot operate as assumed, so Clausius statement must also be correct.

🔍 Proof by reductio ad absurdum (Kelvin–Planck → Clausius)

The reverse direction:

Step 1: Assume the Kelvin–Planck statement is incorrect (the "Kelvin–Planck engine" is possible).

Step 2: Combine the Kelvin–Planck engine with a Carnot heat pump:

  • Make work performed on heat pump equal to work performed by engine in one cycle.

Step 3: One cycle of the combined system produces a device that the Clausius statement says is impossible.

Conclusion: If Clausius is correct, then Kelvin–Planck must also be correct.

Overall result: The truth of one statement implies the truth of the other—they are logically equivalent and either can serve as the fundamental principle for deriving the mathematical statement of the second law.

⚙️ Efficiency of a Carnot engine

📐 Efficiency definition and formula

Efficiency (η): the fraction of the heat input q_h that is returned as net work done on the surroundings.

The defining equation:

  • η = −w / q_h (definition)

From the first law integrated over one cycle:

  • 0 = q_h + q_c + w (one cycle, since internal energy U returns to initial value)
  • Therefore: w = −q_h − q_c

Substituting into efficiency:

  • η = 1 + q_c / q_h (Carnot engine)

📊 Understanding efficiency values

Key constraints:

  • q_c is negative (heat out to cold reservoir).
  • q_h is positive (heat in from hot reservoir).
  • q_c is smaller in magnitude than q_h.
  • Therefore efficiency is less than one (cannot be 100%).

Example: Figure 4.5(a) shows a Carnot engine with η = 1/4 (25% efficiency)—4 joules in, 1 joule of work out, 3 joules exhausted to cold reservoir.

🔬 Comparing efficiencies between Carnot engines

The excerpt begins exploring whether two Carnot engines operating between the same temperatures T_h and T_c could have different efficiencies:

  • Consider a Carnot engine combined with a Carnot heat pump (both between same temperatures).
  • Adjust cycles to produce zero net work for the combined supersystem.
  • The question posed: "Could the efficiency of the Carnot engine be different from the efficiency the heat pump would have when run in reverse as a Carnot engine?"

Don't confuse: The excerpt sets up this question but does not complete the answer—this is leading to an important conclusion about Carnot efficiency depending only on temperatures, not on working substance or design details.

33

Expressions for Heat Capacity

5.6 Expressions for Heat Capacity

🧭 Overview

🧠 One-sentence thesis

All Carnot engines operating between the same two temperatures have the same efficiency, which depends only on the temperature ratio and establishes the thermodynamic temperature scale through the relationship between heat transfers and temperature.

📌 Key points (3–5)

  • Efficiency definition: the fraction of heat input that is converted to net work; for a Carnot engine it equals 1 minus the ratio of cold to hot reservoir temperature.
  • Universal efficiency: all Carnot engines between the same two temperatures have identical efficiency, regardless of the working substance.
  • Thermodynamic temperature: defined by the ratio of heat transfers in a Carnot cycle, making temperature measurable through reversible engine behavior.
  • Common confusion: Carnot engines are idealized limiting processes that real systems approach but never reach; statements about reversible processes describe theoretical limits, not actual operations.
  • Equivalence of temperature scales: the ideal-gas temperature scale is proportional to (and in kelvins, identical to) the thermodynamic temperature scale.

🔧 Carnot engine efficiency fundamentals

🔧 What efficiency measures

Efficiency (η): the fraction of the heat input that is returned as net work done on the surroundings.

  • Mathematically: efficiency equals negative work divided by heat input (η = −w/qₕ).
  • For one complete cycle of a Carnot engine, the first law gives: 0 = qₕ + qc + w (where qₕ is positive heat input, qc is negative heat output, and w is negative work done on surroundings).
  • Rearranging: η = 1 + qc/qₕ.
  • Because qc is negative and smaller in magnitude than qₕ, efficiency is always less than one.
  • Example: the excerpt shows a Carnot engine with η = 1/4, meaning only one-quarter of the heat input becomes work.

🔧 Why efficiency is always less than unity

  • The Kelvin–Planck statement of the second law: a heat engine cannot in one cycle convert all energy transferred by heat from a single heat reservoir into work.
  • Some heat must always be rejected to the cold reservoir (qc is always nonzero and negative).
  • This is a fundamental physical limitation, not an engineering problem to be solved.

🎯 The universal efficiency result

🎯 All Carnot engines have the same efficiency

  • Consider combining a Carnot engine with a Carnot heat pump (the engine run in reverse) operating between the same two temperatures Tₕ and Tc.
  • Adjust the cycles so the combined system (supersystem) produces zero net work for one complete cycle.
  • If the engine and heat pump had different efficiencies, the supersystem would either:
    • Transfer heat from cold to hot with no work input (impossible by the Clausius statement), or
    • When run in reverse, create the same impossibility.
  • Conclusion: all Carnot engines operating between the same two temperatures have the same efficiency.

🎯 Efficiency depends only on temperature

  • Since all Carnot engines between the same two temperatures have identical efficiency, efficiency cannot depend on the working substance (water, air, any gas, etc.).
  • The ratio qc/qₕ must be a unique function of only Tc and Tₕ.
  • This allows us to choose any convenient working substance (like an ideal gas) to derive the functional form.

⚠️ Understanding the reversible limit

  • Carnot engine steps are reversible changes, which cannot actually occur in real systems.
  • Don't confuse: "a Carnot engine operates" does not mean a real device exists.
  • Correct interpretation: all real heat engines have a limiting efficiency approached as the cycle steps are carried out more and more slowly (approaching reversibility).
  • Any irreversibility (friction, finite temperature differences, turbulence) causes efficiency to be lower than the theoretical Carnot value.

🧮 Deriving the efficiency formula

🧮 Using an ideal gas as working substance

  • For an ideal gas: pV = nRT (equation of state).
  • Internal energy change: dU = Cᵥ dT (where Cᵥ is heat capacity at constant volume, a function only of T).
  • Reversible expansion work: đw = −(nRT/V) dV.
  • First law: đq = Cᵥ dT + (nRT/V) dV.
  • Dividing by T: đq/T = Cᵥ (dT/T) + nR (dV/V).

🧮 Heat transfers in isothermal steps

  • In the two isothermal steps (constant temperature), dT = 0.
  • Path A→B (hot reservoir, temperature Tₕ): qₕ = nRTₕ ln(Vʙ/Vᴀ).
  • Path C→D (cold reservoir, temperature Tc): qc = nRTc ln(Vᴅ/Vc).
  • The ratio: qc/qₕ = (Tc/Tₕ) × [ln(Vᴅ/Vc) / ln(Vʙ/Vᴀ)].

🧮 Relating volumes through adiabatic steps

  • In the two adiabatic steps (no heat transfer), đq = 0.
  • Integrating đq/T = 0 over paths B→C and D→A gives volume relationships.
  • Result: ln(Vʙ/Vᴀ) = −ln(Vᴅ/Vc).
  • This simplifies the heat ratio to: qc/qₕ = −Tc/Tₕ.

🧮 Final efficiency expression

  • Substituting into η = 1 + qc/qₕ gives: η = 1 − Tc/Tₕ.
  • Here Tc and Tₕ are temperatures on the ideal-gas scale.
  • But since this result must hold for any working substance, it is universally valid.
  • Example: if η = 1/4, then Tc/Tₕ = 3/4 (e.g., Tc = 300 K and Tₕ = 400 K).

🌡️ Thermodynamic temperature scale

🌡️ Definition through Carnot cycles

Thermodynamic temperature: defined by the relation Tc/Tₕ = −qc/qₕ for a Carnot cycle.

  • The ratio of thermodynamic temperatures of two heat reservoirs equals the ratio of absolute quantities of heat transferred in the isothermal steps.
  • This definition was proposed by Kelvin in 1848 to establish an "absolute" temperature scale.
  • In principle: measure qc/qₕ during a Carnot cycle, use a defined reference temperature (triple point of H₂O = 273.16 kelvins exactly), and calculate the other temperature.

🌡️ Theoretical vs practical measurement

  • A reversible Carnot engine as a "thermometer" is only a theoretical concept.
  • Completely reversible processes cannot occur in practice.
  • The thermodynamic temperature scale is established by the behavior of a heat engine in the reversible limit.

🌡️ Equivalence with ideal-gas scale

  • Both Equation 4.3.13 (ideal-gas temperatures) and Equation 4.3.15 (thermodynamic temperatures) equate Tc/Tₕ to −qc/qₕ.
  • This means: the ratio of ideal-gas temperatures equals the ratio of thermodynamic temperatures for the same two bodies.
  • Therefore the two scales are proportional to one another.
  • The proportionality factor is unity if the same unit (kelvins) is used in both scales.
  • Conclusion: the ideal-gas temperature scale expressed in kelvins is identical to the thermodynamic temperature scale in kelvins.

📐 Historical context

📐 Clausius and the laws of the universe

  • Clausius introduced the concept of entropy for the entire universe.
  • He expressed two fundamental laws:
    1. The energy of the universe is constant.
    2. The entropy of the universe tends to a maximum.
  • These correspond to the first and second laws of thermodynamics applied universally.

📐 Clausius's personal history

  • Clausius was a patriotic German.
  • During the Franco-Prussian war (1870–71), he led an ambulance corps of Bonn students.
  • He was wounded in the leg during battles and suffered disability for the rest of his life.
34

Carnot Engine Efficiency and Thermodynamic Temperature

5.7 Surface Work

🧭 Overview

🧠 One-sentence thesis

The efficiency of a Carnot engine depends only on the temperature ratio of the two heat reservoirs, and this ratio defines the thermodynamic temperature scale, which is identical to the ideal-gas temperature scale when both are expressed in kelvins.

📌 Key points (3–5)

  • Carnot efficiency formula: efficiency equals 1 minus the ratio T_c/T_h, where T_c and T_h are the cold and hot reservoir temperatures.
  • Kelvin–Planck statement: no heat engine can have 100% efficiency—it cannot convert all heat from a single reservoir into work in one cycle.
  • Thermodynamic temperature definition: the ratio of thermodynamic temperatures of two reservoirs equals the ratio of absolute heat quantities transferred in a Carnot cycle's isothermal steps.
  • Common confusion: ideal-gas temperature vs thermodynamic temperature—both scales are proportional and identical when expressed in kelvins, but they are defined differently (one by gas behavior at zero pressure, the other by reversible engine behavior).
  • Irreversibility lowers efficiency: any irreversible step in a real cycle causes dissipation and reduces efficiency below the theoretical Carnot value.

🔥 Carnot engine efficiency

🔥 The efficiency equation

Carnot engine efficiency: η = 1 − (T_c / T_h)

  • What it means: efficiency depends only on the temperature ratio of the cold and hot reservoirs.
  • The ratio T_c/T_h is positive but less than one, so efficiency is always less than one.
  • Why it matters: this is the maximum theoretical efficiency for any heat engine operating between two temperatures.

📉 Efficiency is always less than unity

  • The Kelvin–Planck statement of the second law says a heat engine cannot have efficiency of unity.
  • In plain language: you cannot convert all heat from a single reservoir into work in one cycle.
  • Example: if efficiency η = 1/4, then T_c/T_h = 3/4 (e.g., T_c = 300 K and T_h = 400 K).

⚠️ Irreversibility reduces efficiency

  • The Carnot efficiency formula applies only to a reversible engine operating between two heat reservoirs.
  • If any part of the cycle is irreversible: dissipation of mechanical energy causes the actual efficiency to be lower than the theoretical value.
  • Don't confuse: the formula gives the upper limit; real engines always perform worse.

🌡️ Thermodynamic temperature scale

🌡️ Definition by heat ratios

Thermodynamic temperature: T_c / T_h = −(q_c / q_h) for a Carnot cycle

  • How it works: the ratio of thermodynamic temperatures of two reservoirs is defined as equal to the ratio of the absolute quantities of heat transferred in the isothermal steps of a Carnot cycle.
  • The negative sign accounts for the direction of heat flow (q_c is heat rejected to the cold reservoir).
  • Kelvin's proposal (1848): use this ratio to establish an "absolute" temperature scale that does not depend on any particular substance.

📏 Establishing the scale

  • In principle: measure q_c/q_h during a Carnot cycle, combine with a defined reference temperature, and you can determine the thermodynamic temperature of the other reservoir.
  • The reference point: the triple point of H₂O is defined as exactly 273.16 kelvins.
  • Important caveat: a reversible Carnot engine as a "thermometer" is only a theoretical concept, not a practical instrument, because completely reversible processes cannot occur in practice.

🔗 Relationship between temperature scales

🔗 Ideal-gas vs thermodynamic temperature

ScaleHow it's definedKey equation
Ideal-gasGas thermometer measurements in the limit of zero pressureT_c / T_h = −(q_c / q_h) with T referring to ideal-gas temperatures
ThermodynamicHeat ratios in a reversible Carnot cycleT_c / T_h = −(q_c / q_h) with T referring to thermodynamic temperatures
  • Both equations equate the same ratio: T_c/T_h = −(q_c/q_h)
  • In one equation, T refers to ideal-gas temperatures; in the other, T refers to thermodynamic temperatures.
  • Conclusion: the ratio of ideal-gas temperatures equals the ratio of thermodynamic temperatures for the same two bodies.

✅ The scales are identical

  • Because the ratios are equal, the two scales are proportional to one another.
  • The proportionality factor is arbitrary, but must be unity if the same unit (e.g., kelvins) is used in both scales.
  • Result: the ideal-gas temperature scale and the thermodynamic temperature scale, both expressed in kelvins, are identical.
  • Don't confuse: they are defined by different physical principles (gas behavior vs engine behavior) but yield the same numerical values.

🧮 The Clausius inequality setup

🧮 Experimental system and supersystem

  • The excerpt begins deriving the Clausius inequality for an arbitrary cyclic process of a closed system (called the "experimental system").
  • No restrictions on the experimental system: it may have any complexity, involve multiple kinds of work, phase changes, reactions, temperature and pressure gradients, constraints, external fields, etc.
  • Process requirements: all parts must be either reversible or irreversible, but not impossible.

🔄 The special procedure

  • The experimental cycle is carried out in a special way: heat transferred across the boundary is exchanged with a hypothetical Carnot engine.
  • The supersystem: the combination of the experimental system and the Carnot engine, which is closed.
  • Single heat reservoir: the supersystem exchanges heat with only one heat reservoir of arbitrary constant temperature T_res.
  • Why this setup: by using a single reservoir, the Kelvin–Planck statement can be applied to a cycle of the supersystem.

🎛️ Control from the surroundings

  • Changes to the work coordinates of the experimental system are controlled from the surroundings of the supersystem.
  • The Carnot engine is also controlled from these surroundings (e.g., by moving a piston).
  • Result: energy transferred by work across the experimental system boundary and work required to operate the Carnot engine are both exchanged with the surroundings of the supersystem.
  • During each stage with nonzero heat, the Carnot engine undergoes many infinitesimal Carnot cycles with infinitesimal quantities of heat and work.
35

Criteria for Spontaneity

5.8 Criteria for Spontaneity

🧭 Overview

🧠 One-sentence thesis

The second law of thermodynamics establishes that entropy is a state function whose change during any process in a closed system equals ∂q/T_b for reversible processes and exceeds ∂q/T_b for irreversible (spontaneous) processes, providing a mathematical criterion to distinguish possible from impossible changes.

📌 Key points (3–5)

  • Entropy as a state function: Entropy S is defined so that dS = ∂q_rev/T_b for reversible processes; its value depends only on the equilibrium state, not on the path taken to reach it.
  • The fundamental inequality: For any infinitesimal change in a closed system, dS ≥ ∂q/T_b, with equality for reversible processes and strict inequality for irreversible processes.
  • Spontaneity in isolated systems: In an isolated system, entropy continuously increases during any spontaneous (irreversible) process until it reaches a maximum at equilibrium (dS > 0).
  • Common confusion—path vs. state: The value of ΔS between two states is always the same (state function), but the integral of ∂q/T_b depends on the path; for irreversible paths, the integral is less than ΔS.
  • Why it matters: These criteria allow prediction of whether a process can occur spontaneously, determine equilibrium conditions, and explain the direction of natural processes like heat flow and gas expansion.

🔬 Deriving entropy from reversible processes

🔬 Reversible adiabatic surfaces

The excerpt establishes that equilibrium states can be organized into families of "reversible adiabatic surfaces" in an N-dimensional state space (where N is the number of independent variables).

  • What they are: Each surface contains all equilibrium states that can be reached from one another by reversible adiabatic processes (no heat transfer).
  • Key property: These surfaces never intersect—they fill the entire state space without touching.
  • Why they exist: Carathéodory's principle of adiabatic inaccessibility states that every equilibrium state has other states infinitesimally close to it that cannot be reached by reversible adiabatic processes.
  • Example: If you heat a liquid reversibly (adding heat), you move from one surface to another; you cannot duplicate this temperature change using only reversible work.

📐 Defining entropy on these surfaces

Entropy S: a state function assigned constant values on each reversible adiabatic surface, with different unique values on different surfaces.

  • The existence of non-intersecting reversible adiabatic surfaces guarantees that entropy can be defined as a state function.
  • Assignment rule: A reversible process with positive heat corresponds to increasing S; negative heat corresponds to decreasing S.
  • The formula: For a reversible process changing state A to state B, ΔS = integral of (∂q_rev/T_b).
  • This integral has the same value for any reversible path between the same two surfaces, making it a valid state function definition.

🔄 The Clausius inequality foundation

The derivation uses a supersystem combining the experimental system with a Carnot engine and heat reservoir.

  • Setup: The Carnot engine mediates heat transfer, operating many infinitesimal cycles while the system undergoes its process.
  • Key result: For any cyclic process of a closed system, the cyclic integral of (∂q/T_b) ≤ 0 (the Clausius inequality).
  • For a reversible cycle specifically: the cyclic integral equals exactly zero.
  • Implication: This proves that ∂q_rev/T_b is an exact differential (its integral depends only on endpoints, not path).

⚖️ Reversible vs. irreversible processes

⚖️ The equality for reversible changes

For any infinitesimal reversible change in a closed system:

dS = ∂q_rev/T_b

  • T_b is the temperature at the boundary where heat is transferred.
  • If temperature is uniform throughout the system, this becomes dS = ∂q_rev/T.
  • Integration gives: ΔS = integral of (∂q_rev/T_b) for the entire reversible process.

⚖️ The inequality for irreversible changes

For any infinitesimal irreversible change in a closed system:

dS > ∂q/T_b

  • The entropy change is strictly greater than the heat divided by boundary temperature.
  • Why: The derivation shows that in an irreversible adiabatic process, dS must be positive (cannot be zero or negative), and this extends to all irreversible processes through the supersystem construction.
  • Integration gives: ΔS > integral of (∂q/T_b) for an irreversible process.

🔀 The crucial distinction

AspectReversible pathIrreversible path
Entropy change ΔSSame value (state function)Same value (state function)
Integral of ∂q/T_bEquals ΔSLess than ΔS
Process characterContinuous equilibrium statesIncludes non-equilibrium states

Don't confuse: ΔS is always the same between two given states regardless of path; what differs is the value of the integral ∫(∂q/T_b), which equals ΔS only for reversible paths.

🎯 Applications and spontaneity criteria

🎯 Isolated systems—the master criterion

For an isolated system (no exchange of matter or energy with surroundings):

dS > 0 (for any irreversible/spontaneous change)

  • The principle: Entropy continuously increases during spontaneous processes until reaching a maximum at equilibrium.
  • Any spontaneous change in an isolated system must increase entropy.
  • At equilibrium, entropy is at its maximum possible value for the constraints.
  • Example: A gas expanding into a vacuum (free expansion) is isolated, irreversible, and has positive ΔS even though q = 0 and w = 0.

🌡️ Internal heat flow

When a thermally-insulated body has non-uniform temperature initially:

  • Heat flows spontaneously from warmer to cooler regions.
  • Entropy calculation: Treat the body as many small phases, each at temperature T_φ.
  • For heat ∂q_φψ flowing from phase ψ to phase φ: the contribution to dS is (1/T_φ - 1/T_ψ)∂q_φψ.
  • This is always positive when T_φ < T_ψ (heat flows to cooler phase).
  • Result: Entropy increases continuously until uniform temperature is reached.
  • Don't confuse: Even though the system is isolated (adiabatic boundary), internal heat transfer occurs and drives entropy increase.

🔧 Reversible heating and expansion

Reversible heating (constant volume or pressure):

ΔS = integral from T₁ to T₂ of (C/T)dT

  • If heat capacity C is constant: ΔS = C ln(T₂/T₁)
  • Heating increases entropy; cooling decreases it.

Reversible isothermal expansion of ideal gas:

ΔS = nR ln(V₂/V₁)

  • Expansion (V₂ > V₁) increases entropy; compression decreases it.
  • This formula is valid whenever T₂ = T₁, even if intermediate states are not equilibrium states.

⚙️ Adiabatic processes with work

For an irreversible adiabatic process with work:

  • General principle: The work w_irr is algebraically greater than the reversible work w_rev for the same change of state.
  • To achieve the same final state reversibly: (1) perform reversible adiabatic work w_rev, then (2) add positive heat q_rev.
  • Since q_rev > 0, the entropy change is positive.
  • Conclusion: Irreversible adiabatic processes with work always increase entropy, consistent with dS > ∂q/T_b = 0.

📊 The combined mathematical statement

📊 The general relation

For any infinitesimal change in a closed system:

dS ≥ ∂q/T_b

  • The inequality (>) applies to irreversible changes.
  • The equality (=) applies to reversible changes.
  • Integrated form: ΔS ≥ integral of (∂q/T_b).

📊 Conditions and limitations

Valid for:

  • Closed systems (no matter exchange)
  • Both reversible and irreversible processes
  • Any combination of heat and work

Not valid for:

  • Open systems (matter can enter/leave; entropy can increase or decrease)
  • Impossible processes (no general relation established)

📊 Entropy as an extensive property

  • For a system divided into subsystems A and B at uniform temperature T: S = S_A + S_B.
  • Entropy changes are additive: dS = dS_A + dS_B.
  • This confirms entropy is an extensive state function.

🔬 Statistical interpretation

🔬 The microscopic meaning

Statistical entropy: S_stat = k ln(W) + C, where k is Boltzmann's constant, W is the number of accessible microstates, and C is a constant.

  • Accessible microstates: quantum states compatible with the system's constraints and energy.
  • Fundamental assumption: All accessible microstates of equal energy are equally probable.
  • The system continually jumps between microstates while in a macroscopic equilibrium state.

🔬 Entropy increase as energy dispersal

  • When entropy increases in a reversible process, the number of accessible microstates W increases.
  • Better interpretation than "disorder": Entropy increase represents the spreading and dispersal of energy among more microstates.
  • This explains reversible heating (more energy states become accessible) and isothermal expansion (more spatial configurations available).

🔬 Why not "disorder"?

Don't confuse entropy with disorder:

  • The "disorder" interpretation fails to explain reversible heating at constant volume.
  • It seems contradicted by freezing of supercooled liquids (spontaneous in isolated system, but appears to decrease disorder).
  • The statistical interpretation as energy dispersal is more rigorous and consistent.
36

The Zero of Entropy

6.1 The Zero of Entropy

🧭 Overview

🧠 One-sentence thesis

The third law establishes that every pure, perfectly-ordered crystalline substance has zero entropy at absolute zero temperature, creating a universal reference point for measuring absolute entropies at higher temperatures.

📌 Key points (3–5)

  • The Nernst heat theorem (third law): entropy change approaches zero as temperature approaches zero kelvins for pure, perfectly-ordered crystals.
  • Arbitrary zero convention: we can choose the entropy of every pure crystalline element to be zero at 0 K; the third law then requires that every pure crystalline compound also has zero entropy at 0 K.
  • Third-law entropies: this convention establishes a scale of absolute entropy values (called third-law entropies) at temperatures above zero kelvins.
  • Common confusion: the third law applies only to pure crystals with perfect spatial order—imperfect crystals or mixtures do not necessarily have zero entropy at 0 K.
  • Why it matters: absolute entropy values can be tabulated and used in calculations for chemical reactions and phase transitions.

🧊 The third law statement

🧊 Mathematical form of the Nernst heat theorem

The third law of thermodynamics (Nernst heat theorem): the limit as temperature approaches zero of the entropy change ΔS equals zero, for pure, perfectly-ordered crystals.

  • Written as: limit as T approaches 0 of ΔS = 0 (for pure, perfectly-ordered crystals).
  • This applies to entropy changes in chemical reactions or phase transitions studied at low temperatures.
  • Key restriction: all substances involved must be pure crystals with identical unit cells arranged in perfect spatial order.

🔬 Historical context

  • Nernst originally formulated his heat theorem in 1906 using a different function (the partial derivative of Helmholtz energy with respect to temperature at constant volume), avoiding the entropy function.
  • The modern statement uses entropy directly because it is more general and easier to apply.

🎯 Establishing the zero point

🎯 Why we can choose zero arbitrarily

  • No theoretical relation exists between the entropies of different chemical elements.
  • We are free to set the entropy of every pure crystalline element to zero at zero kelvins as a convention.
  • This is an arbitrary choice, not a physical requirement—it simply establishes a reference scale.

🔗 How the convention extends to compounds

  • Once we set all pure crystalline elements to zero entropy at 0 K, the third law (Equation 6.0.1) requires that every pure crystalline compound also has zero entropy at 0 K.
  • Why: Consider forming a compound from its elements at 0 K. The third law says the entropy change for this reaction approaches zero as T approaches 0 K. If the elements start at zero entropy, the compound must also end at zero entropy.
  • Example: If we form a compound from Sender-element and Receiver-element at 0 K, and both elements have zero entropy, the compound must also have zero entropy for the entropy change to be zero.

📜 The Lewis and Randall principle (1923)

"If the entropy of each element in some crystalline state be taken as zero at the absolute zero of temperature: every substance has a finite positive entropy, but at the absolute zero of temperature the entropy may become zero, and does so become in the case of perfect crystalline substances."

  • This classic statement emphasizes two points:
    1. At temperatures above 0 K, every substance has finite positive entropy.
    2. At 0 K, entropy becomes zero only for perfect crystalline substances.
  • Don't confuse: "may become zero" vs. "does become zero"—the entropy does become zero for perfect crystals, but may not for imperfect or disordered substances.

📐 The entropy scale convention

📐 Molar entropy at absolute zero

  • For any pure, perfectly-ordered crystal at 0 K and any pressure, the molar entropy is defined as zero: S_m(0 K) = 0.
  • Pressure independence: The excerpt notes (in a footnote) that entropy becomes independent of pressure as temperature approaches zero kelvins.

📊 Third-law entropies

  • Definition: Absolute entropy values measured from the zero point established by the third law convention.
  • These are the entropy values typically tabulated for use in calculations.
  • The next section (6.2) explains how to calculate third-law molar entropies at temperatures above 0 K by integrating heat capacity and accounting for phase transitions.

⚠️ Critical conditions and limitations

⚠️ Perfect order requirement

ConditionThird law applies?Reason
Pure, perfectly-ordered crystalYesIdentical unit cells in perfect spatial order
Imperfect crystal (defects, disorder)NoResidual entropy may remain at 0 K
Mixture or solutionNoConfigurational entropy may remain at 0 K
  • The third law is "true in general only if each reactant and product is a pure crystal with identical unit cells arranged in perfect spatial order."
  • Common confusion: Not all solids at 0 K have zero entropy—only those with perfect crystalline order do.

🔍 What "perfect spatial order" means

  • Identical unit cells: every repeating unit of the crystal structure is the same.
  • Perfect arrangement: no defects, vacancies, or disorder in how the unit cells are positioned.
  • Example: A crystal with random orientations of molecules in different unit cells would not qualify as perfectly ordered, even at 0 K.
37

Molar Entropies

6.2 Molar Entropies

🧭 Overview

🧠 One-sentence thesis

The Third Law convention that perfectly-ordered crystals have zero entropy at 0 K establishes an absolute scale for calculating molar entropies at any temperature through calorimetric measurements or spectroscopic theory, though some substances show residual entropy due to frozen-in disorder.

📌 Key points (3–5)

  • Third-law convention: Pure, perfectly-ordered crystals at 0 K have zero molar entropy, establishing an absolute entropy scale.
  • How to calculate molar entropy: Integrate heat capacity divided by temperature from 0 K to the desired temperature, adding entropy changes from phase transitions.
  • Two methods: Calorimetric (third-law) method uses heat capacity measurements; spectroscopic method uses statistical mechanics and molecular properties for ideal gases.
  • Residual entropy: Discrepancies between calorimetric and spectroscopic values reveal frozen-in disorder in crystals that never achieved perfect order.
  • Common confusion: The practical entropy scale ignores isotope mixing and nuclear spin randomness above 1 K, but this causes no error in entropy change calculations because these effects cancel out.

🧊 The Third Law foundation

🧊 Zero entropy convention

The molar entropy of a pure, perfectly-ordered crystal at 0 K, at any pressure, is zero: S_m(0 K) = 0

  • This convention sets the reference point for all entropy measurements.
  • The pressure independence at 0 K comes from experimental observation that the cubic expansion coefficient approaches zero as temperature approaches zero.
  • This establishes what are called third-law entropies—absolute entropy values at temperatures above zero kelvins.

📏 Why absolute entropies matter

  • Unlike enthalpy, where only changes matter, entropy can be assigned absolute values.
  • These absolute values are what get tabulated for practical calculations.
  • The zero-point convention makes it possible to calculate entropy at any state by tracking the path from 0 K.

🔬 Calorimetric method for molar entropy

🔬 The heating path from 0 K

To find the molar entropy at temperature T₀ and a given pressure:

  • Start with the perfectly-ordered crystal at 0 K (entropy = 0).
  • Reversibly heat at constant pressure to the desired state.
  • The entropy change of this process equals the molar entropy at T₀.

Key insight: Since entropy is a state function, any reversible path gives the same result.

🧮 The calculation formula

The operational equation for molar entropy is:

S_m(T₀) = integral from 0 to T₀ of (C_p,m / T) dT + sum of (ΔH_trs / T_trs)

Where:

  • C_p,m is the molar heat capacity at constant pressure
  • The sum covers all equilibrium phase transitions during heating
  • ΔH_trs is the molar enthalpy of each transition
  • T_trs is the temperature of each transition

Why both terms?

  • The integral accounts for continuous heating within a single phase.
  • The sum accounts for discontinuous jumps at phase transitions (solid→solid, melting, vaporization).

❄️ Low-temperature approximation

The measurement problem: Calorimeters can only measure down to about 10 K using liquid hydrogen cooling.

The solution: Use Debye theory from statistical mechanics:

  • For nonmagnetic nonmetals: C_p,m is proportional to T³ at low temperatures (0 K to ~30 K).
  • This allows extrapolation: the contribution from 0 K to the lowest measured temperature T'' is approximately C_p,m(T'') / 3.

The practical formula becomes:

S_m(T₀) = C_p,m(T'') / 3 + integral from T'' to T₀ of (C_p,m / T) dT + sum of (ΔH_trs / T_trs)

For metals: An additional electronic term proportional to T should be added, but the error from ignoring it is usually negligible if measurements go down to ~10 K.

📊 Numerical integration

Two equivalent ways to evaluate the integral:

  1. Plot C_p,m / T versus T and find the area under the curve.
  2. Plot C_p,m versus ln(T/K) and find the area under the curve.

Both give the same numerical result because the integral can be written as: integral of (C_p,m / T) dT = integral of C_p,m d[ln(T/K)]

Example: Figure 6.1 shows hydrogen chloride (HCl) data. The molar entropy increases continuously with temperature and shows discontinuities (jumps) at each phase transition—solid→solid, melting, and vaporization.

✅ Always positive

Since C_p,m is positive at all temperatures above 0 K, and all transition enthalpies during heating are positive, the molar entropy of any substance is positive at all temperatures above 0 K.

🌟 Spectroscopic method for ideal gases

🌟 Statistical mechanics approach

For ideal gases, statistical mechanical theory provides an accurate alternative to calorimetry.

  • Uses experimental molecular properties from spectroscopy.
  • Ignores intermolecular interactions (ideal gas assumption).
  • Uses the same practical entropy scale (ignores isotope mixing and nuclear spin interactions).

Why this method? Often the preferred method for gases because it can be more accurate and doesn't require difficult low-temperature measurements.

🧬 The calculation

Molar entropy is split into two contributions:

S_m = S_m,trans + S_m,int

Translational contribution (Sackur–Tetrode equation): S_m,trans depends on molecular mass M, temperature T, pressure p, and fundamental constants (Planck constant h, Avogadro constant N_A).

Internal contribution: S_m,int depends on the molecular partition function q_int, which is calculated from:

  • Energies of low-lying electronic energy levels
  • Electronic degeneracies
  • Fundamental vibrational frequencies
  • Rotational constants
  • Other spectroscopic parameters

The partition function sums over all quantum states of one molecule, weighted by their Boltzmann factors, with averaging for natural isotopic abundance.

📐 Standard molar entropy

When evaluated at standard pressure p° = 1 bar, the result is the standard molar entropy S°_m.

  • Useful for thermodynamic calculations even if the real gas isn't ideal at 1 bar.
  • Provides a reference value for comparing different substances.

🔍 Residual entropy and crystal disorder

🔍 Comparing the two methods

Ideally, calorimetric and spectroscopic values of S°_m should agree closely.

Table 6.1 comparison at 298.15 K and 1 bar:

SubstanceCalorimetric S°_mSpectroscopic S°_mResidual entropy S_m,0
HCl186.3 ± 0.4186.9010.6 ± 0.4
CO193.4 ± 0.4197.65 ± 0.044.3 ± 0.4
NO208.0 ± 0.4210.7582.8 ± 0.4
N₂O215.3 ± 0.4219.9574.7 ± 0.4
H₂O185.4 ± 0.2188.834 ± 0.0423.4 ± 0.2

(All entropy values in J K⁻¹ mol⁻¹)

Residual entropy S_m,0 = (spectroscopic value) − (calorimetric value)

🧩 What residual entropy means

The spectroscopic value is the true entropy because:

  • It assumes the solid has only one microstate at 0 K (entropy = 0).
  • It accounts for all accessible microstates of the ideal gas at 298.15 K.

The calorimetric value assumes perfect ordering as T → 0 K, which may not happen.

Why HCl shows good agreement: Polar, asymmetric molecules have a large energetic advantage for forming perfectly-ordered crystals, so the assumption holds.

Why others show residual entropy: Almost-symmetric molecules with small dipole moments can have random rotational orientations of nearly equal energy in the crystal.

🔄 Frozen-in disorder

The conventional explanation:

  • Random molecular orientations are established when crystals form from the liquid.
  • This randomness becomes frozen into the crystals as temperature drops below the freezing point.
  • The crystal never achieves perfect order, even at the lowest measurable temperatures.

Example: H₂O (water): The residual entropy comes from random arrangement of intermolecular hydrogen bonds, not molecular orientations.

Other contributors: Crystal imperfections such as dislocations can also add to residual entropy.

🎯 Physical interpretation

If crystal imperfection is present at the lowest experimental temperature:

  • The calorimetric S°_m at 298.15 K represents the entropy increase from the imperfectly-ordered solid at 0 K to the ideal gas at 298.15 K.
  • The residual entropy S_m,0 is the molar entropy of this imperfectly-ordered solid at 0 K.

Don't confuse: This is not a failure of the Third Law. The Third Law states that perfectly-ordered crystals have zero entropy at 0 K. Residual entropy reveals that some substances never achieve perfect order.

🛠️ The practical entropy scale

🛠️ Two departures from the ideal

The experimental entropy scale differs from the strict Third Law convention in two ways:

  1. Isotopic mixing: Elements are usually mixtures of isotopes, not isotopically pure.
  2. Nuclear spin randomness: Above ~1 K, nuclear spin orientations are essentially random. Below 1 K, weak interactions would cause ordering, but this is not included in the Debye T³ formula.

✨ Why this doesn't matter

The practical (conventional) entropy scale assigns zero entropy to a crystal that:

  • Has randomly-mixed isotopes
  • Has randomly-oriented nuclear spins
  • Is pure and ordered in all other respects

No inaccuracies result for any process occurring above 1 K because:

  • The shift in the zero point is the same in initial and final states.
  • Isotopes remain randomly mixed throughout.
  • Nuclear spins remain randomly oriented throughout.
  • These contributions cancel when calculating ΔS for any process.

Key insight: For entropy changes (which is what we use in thermodynamics), the choice of zero point doesn't matter as long as it's consistent.

38

Cryogenics

6.3 Cryogenics

🧭 Overview

🧠 One-sentence thesis

Cryogenic techniques—including Joule–Thomson expansion and adiabatic demagnetization—enable the production of extremely low temperatures needed for calorimetric entropy measurements and demonstrate that absolute zero cannot be reached in a finite number of steps.

📌 Key points (3–5)

  • Joule–Thomson expansion: a throttling process that cools most gases by forcing them through a constriction at constant enthalpy, though hydrogen and helium require pre-cooling to work.
  • Adiabatic demagnetization: a technique using paramagnetic solids that achieves temperatures below 1 K by first magnetizing isothermally (removing entropy) then demagnetizing adiabatically (lowering temperature).
  • Common confusion: the Joule–Thomson coefficient can be positive or negative depending on the gas and temperature—hydrogen and helium have negative coefficients at room temperature, so throttling would heat them rather than cool them.
  • Unattainability of absolute zero: because entropy curves converge at 0 K, no finite sequence of adiabatic demagnetization cycles can reach absolute zero.
  • Why cryogenics matters: very low temperatures are essential for evaluating third-law entropies through calorimetric measurements and studying matter's behavior at extreme conditions.

🌡️ Joule–Thomson expansion

🔧 What the process is

Throttling process: a continuous adiabatic procedure in which a fluid is forced to flow through a porous plug, valve, or other constriction causing an abrupt pressure drop.

  • Also called the Joule–Thomson experiment or Joule–Kelvin experiment (after collaborators working 1852–1862).
  • The gas flows slowly from high pressure (p′) to low pressure (p″) through a porous plug in a thermally insulated tube.
  • A steady state develops: uniform temperature T′ on the high-pressure side, uniform temperature T″ on the low-pressure side.

🔥 Why it is isenthalpic

  • Consider a fixed segment of gas moving through the plug as a closed system.
  • Heat transfer q = 0 (insulated walls, no temperature gradient at the ends).
  • Work done on the segment: w = p′(V′₂ − V′₁) − p″(V″₂ − V″₁).
  • Internal energy change ΔU = w (since q = 0).
  • Enthalpy change ΔH = ΔU + (p′V′₂ + p″V″₂) − (p′V′₁ + p″V″₁) = 0 when combined with the work expression.
  • Result: the gas has the same enthalpy before and after passing through the plug.

📉 The Joule–Thomson coefficient

Joule–Thomson coefficient (or Joule–Kelvin coefficient): μ_JT = (∂T/∂p)_H, the slope of temperature versus pressure at constant enthalpy.

  • For an ideal gas, μ_JT = 0 because enthalpy depends only on T; if H is constant, T cannot change.
  • For a nonideal gas, μ_JT is a function of T, p, and the kind of gas.
Gas typeConditionsSign of μ_JTEffect of expansion
Most gasesLow to moderate pressure, near room temperaturePositiveCooling (T decreases as p decreases)
HydrogenRoom temperatureNegativeHeating; must cool to ~200 K first
HeliumRoom temperatureNegativeHeating; must cool to ~40 K first
  • Don't confuse: the same gas can have different signs of μ_JT at different temperatures—hydrogen is "wrong" at room temperature but "right" at 77.4 K.

❄️ Cascading to liquefy gases

  • Starting at room temperature, nitrogen can be condensed to liquid nitrogen at 77.4 K by throttling.
  • Liquid nitrogen cools hydrogen to 77.4 K, where hydrogen's μ_JT becomes positive, allowing throttling to liquefy it at 20.3 K.
  • Liquid hydrogen cools helium to 20.3 K, where helium's μ_JT is positive, enabling throttling to liquid helium at 4.2 K.
  • Further cooling to ~1 K: pump on liquid helium to cause rapid evaporation.
  • Example: a recirculating system pumps gas through the throttle and uses a heat exchanger so the cooler low-pressure gas pre-cools the high-pressure gas.

🧲 Adiabatic demagnetization

🧩 The principle

Adiabatic demagnetization: a technique to reach temperatures below 1 K by first magnetizing a paramagnetic solid isothermally (ordering magnetic dipoles, reducing entropy) then removing the field adiabatically (entropy constant, temperature drops).

  • Suggested independently by Peter Debye (1926) and William Giauque (1927).
  • Requires a paramagnetic solid with unpaired electrons sufficiently separated that at 1 K the magnetic dipole orientations are almost completely random.
  • Common material: gadolinium(III) sulfate octahydrate, Gd₂(SO₄)₃·8H₂O.

🔄 The two-step cycle

Path A: Isothermal magnetization

  • The solid, surrounded by gaseous helium in thermal contact with liquid helium (~1 K), is slowly moved into a strong magnetic field.
  • The field partially orients the magnetic dipoles, reducing entropy.
  • Heat is transferred to the liquid helium, which partially boils away.
  • Entropy change is negative (because magnetic moment m_mag decreases with increasing T at constant B, so (∂S/∂B){T,p} = (∂m_mag/∂T){p,B} < 0).

Path B: Adiabatic demagnetization

  • Thermal contact is broken by pumping away the surrounding gas.
  • The sample is slowly moved away from the magnetic field.
  • The process is reversible and adiabatic, so entropy change = 0.
  • Because (∂T/∂B)_{S,p} is positive (derived from reciprocity relation and Curie's law), decreasing B decreases T.
  • The solid reaches a lower temperature on the same entropy curve.

📐 Thermodynamic relations

  • Internal energy with magnetization: dU = T dS − p dV + B dm_mag.
  • Magnetic enthalpy: H′ = U + pV − Bm_mag, with total differential dH′ = T dS + V dp − m_mag dB.
  • Reciprocity relation: (∂T/∂B){S,p} = −(∂m_mag/∂S){p,B}.
  • According to Curie's law, m_mag at constant B is proportional to 1/T (or at least decreases with increasing T).
  • To increase T at constant B, heat enters, S increases, so (∂m_mag/∂S)_{p,B} < 0.
  • Therefore (∂T/∂B)_{S,p} > 0: adiabatic demagnetization (B decreases) causes T to decrease.

🏔️ Achieving ultra-low temperatures

  • Repeated cycles of isothermal magnetization and adiabatic demagnetization, starting each stage at the temperature from the previous stage, have reached as low as 0.0015 K.
  • Adiabatic nuclear demagnetization can go even lower: down to 16 microkelvins.
  • Example: starting at 1 K, one cycle might reach 0.25 K; repeating from 0.25 K reaches still lower temperatures.

🚫 Unattainability of absolute zero

🔒 The principle

Principle of the unattainability of absolute zero: it is not possible to reach a temperature of zero kelvins in a finite number of stages of adiabatic demagnetization.

  • The entropy curves for zero field and finite field both converge at 0 K (third law).
  • Each adiabatic demagnetization step is a vertical line (constant entropy) on an S-versus-T diagram.
  • Because the curves come together, the vertical distance (temperature drop) shrinks as you approach 0 K.
  • No finite sequence of steps can bridge the remaining gap to absolute zero.
  • Don't confuse: this is not a practical limitation of equipment—it is a fundamental thermodynamic constraint from the third law.

🎯 Why cryogenics matters

📊 Applications

  • Evaluating third-law entropies: calorimetric measurements of heat capacity down to very low temperatures are needed to calculate entropy via the integral of C_p/T from 0 K to 298.15 K.
  • Studying matter at extreme conditions: behavior of substances at temperatures below 1 K reveals quantum effects and tests fundamental thermodynamic principles.
  • Liquefying gases: industrial and laboratory use of liquid nitrogen (77.4 K), liquid hydrogen (20.3 K), and liquid helium (4.2 K) for cooling and as cryogenic fluids.

🧪 Historical context

  • William Giauque received the 1949 Nobel Prize in Chemistry for contributions to chemical thermodynamics and behavior of substances at extremely low temperatures.
  • His 1933 experiment reported a temperature of 0.25 K using adiabatic demagnetization.
  • Discovery of oxygen isotopes ¹⁷O and ¹⁸O came from studying faint lines in oxygen's absorption spectrum during cryogenic research.
39

Phase Equilibria of Pure Substances

7.1 Volume Properties

🧭 Overview

🧠 One-sentence thesis

A multiphase system of a single substance reaches equilibrium when temperature, pressure, and chemical potential are uniform throughout all phases, and these conditions can be derived systematically by maximizing entropy in an isolated system.

📌 Key points (3–5)

  • Core equilibrium principle: At equilibrium, each phase must have the same temperature (thermal equilibrium), same pressure (mechanical equilibrium), and same chemical potential (transfer equilibrium).
  • Derivation method: Isolate the system, write the entropy differential in terms of independent variables, then require dS = 0 for all infinitesimal changes to find equilibrium conditions.
  • Chemical potential drives transfer: Substance spontaneously transfers from higher to lower chemical potential phases; equilibrium occurs when chemical potentials are equal.
  • Common confusion: In a gravitational field, temperature and chemical potential remain uniform at equilibrium, but pressure varies with elevation—don't assume all three must be uniform in all situations.
  • Why it matters: These conditions predict when phases coexist stably and explain phenomena like pressure variation in tall gas columns and the barometric formula.

🔬 The three equilibrium conditions

🌡️ Thermal equilibrium

Thermal equilibrium: All phases have the same temperature.

  • A temperature difference between phases would cause spontaneous heat transfer from warmer to cooler phase.
  • This transfer would be irreversible and increase entropy until temperatures equalize.
  • Example: If a liquid phase is warmer than its vapor phase, heat flows spontaneously until both reach the same temperature.

⚖️ Mechanical equilibrium

Mechanical equilibrium: All phases have the same pressure.

  • A pressure difference would cause spontaneous flow of matter from higher to lower pressure.
  • This flow continues until pressures equalize.
  • Example: If vapor pressure exceeds liquid pressure, matter flows until pressures match.

⚗️ Transfer equilibrium

Transfer equilibrium: All phases have the same chemical potential.

  • Chemical potential of a pure phase is the Gibbs energy per amount of substance in that phase.
  • Substance spontaneously transfers from higher to lower chemical potential phases.
  • At constant T and p, total Gibbs energy decreases during spontaneous transfer; when chemical potentials are equal, no further spontaneous transfer occurs.
  • Example: If liquid has higher chemical potential than vapor, evaporation occurs spontaneously until chemical potentials equalize.

📐 Systematic derivation procedure

📝 Five-step method

The excerpt presents a general five-step procedure for finding equilibrium conditions:

  1. Write total differential of internal energy U consistent with constraints and independent variables
  2. Impose isolation conditions (dU = 0, no expansion work, closed system), reducing independent variables
  3. Choose reference phase φ₀ and substitute: dS_φ₀ = dS - Σ dS_φ (for all phases φ ≠ φ₀)
  4. Rearrange to obtain dS expression in terms of reduced independent variables
  5. Set dS = 0 for all infinitesimal changes of independent variables to find equilibrium

🧮 Application to multiphase systems

Starting with the differential for each phase (open subsystem with expansion work only):

  • dU = T dS - p dV + μ dn

Isolation constraints:

  • dU = 0 (constant internal energy, adiabatic container)
  • Σ dV_φ = 0 (rigid container, no expansion work)
  • Σ dn_φ = 0 (closed system, no matter exchange with surroundings)

After substitution and rearrangement, the entropy differential becomes:

dS = Σ [(T_φ₀ - T_φ)/T_φ₀] dS_φ - Σ [(p_φ₀ - p_φ)/T_φ₀] dV_φ + Σ [(μ_φ₀ - μ_φ)/T_φ₀] dn_φ

For dS = 0 for all infinitesimal changes, each coefficient must be zero, yielding:

  • T_φ = T_φ₀ for all phases (uniform temperature)
  • p_φ = p_φ₀ for all phases (uniform pressure)
  • μ_φ = μ_φ₀ for all phases (uniform chemical potential)

🎯 Simpler intuitive derivation

The excerpt also provides a less formal approach:

  • Temperature uniformity: Temperature difference causes spontaneous heat transfer; equilibrium requires equal temperatures.
  • Pressure uniformity: Pressure difference causes spontaneous matter flow; equilibrium requires equal pressures.
  • Chemical potential uniformity: At constant T and p, Gibbs energy decreases when substance transfers from higher to lower chemical potential; equilibrium requires equal chemical potentials (no further spontaneous transfer possible).

🌍 Special case: Gravitational field effects

🏔️ Tall gas column equilibrium

When a gas column has considerable vertical extent in Earth's gravitational field:

  • The system is modeled as many thin horizontal slab-shaped phases stacked vertically.
  • Each slab has constant volume (fixed elevation) in the rigid container.
  • The derivation follows the same procedure but with dV_φ = 0 for each phase.

Result: In equilibrium, temperature and chemical potential are uniform throughout, but the derivation gives no information about pressure uniformity.

Don't confuse: Unlike the standard multiphase case, pressure is NOT uniform in a gravitational field—it varies with elevation even at equilibrium.

📉 Fugacity and elevation

The chemical potential at elevation h includes a gravitational term:

μ(h) = μ°(g) + RT ln[f(h)/p°] + Mgh

where:

  • M is molar mass
  • g is gravitational acceleration
  • h is elevation above reference

For chemical potential to be uniform (equilibrium condition), fugacity must decrease with increasing elevation:

f(h) = f(0) exp(-Mgh/RT)

📊 Barometric formula

For an ideal gas (fugacity equals pressure):

p(h) = p(0) exp(-Mgh/RT)

Barometric formula: Pressure in a pure ideal gas at equilibrium decreases exponentially with increasing elevation.

Alternative derivation (shorter, using Newton's second law):

  • Consider a thin slab of gas: area A_s, thickness Δh, density ρ.
  • Vertical forces at rest: upward pressure force pA_s, downward pressure force (p + Δp)A_s, downward gravitational force ρgA_s Δh.
  • Net force = 0 gives: Δp = -ρg Δh
  • In the limit: dp = -ρg dh (general relation for any fluid in gravitational field)
  • For ideal gas, ρ = Mp/RT, leading to the same barometric formula.

🔑 Terminology note

Some thermodynamicists distinguish "total chemical potential" (or "gravitochemical potential") from "chemical potential":

  • "Total chemical potential" = μ°(g) + RT ln(f/p°) + Mgh (uniform at equilibrium)
  • "Chemical potential" = μ°(g) + RT ln(f/p°) (decreases with elevation)

This book's definition: Chemical potential μ is the molar Gibbs energy at a given elevation, which remains uniform at all elevations in equilibrium (following 2001 IUPAC recommendation).

🧪 Physical interpretation

💧 Why chemical potential matters

ConditionWhat happensDirection of change
μ_phase1 > μ_phase2Spontaneous transfer from phase 1 to phase 2Gibbs energy decreases
μ_phase1 = μ_phase2No spontaneous transferEquilibrium (Gibbs energy constant)
  • At constant T and p, when substance transfers between phases, intensive properties (including chemical potential) of each phase remain unchanged.
  • Only the amounts in each phase change.
  • Transfer continues until chemical potentials equalize.

🎈 Gravitational work and chemical potential

When a small gas sample is raised from h = 0 to elevation h:

  • Gravitational work required: w' = mgh = nMgh
  • Process is reversible at constant volume, no heat: T, p, V, S, f unchanged
  • Internal energy U increases by nMgh
  • Gibbs energy G = U - TS + pV also increases by nMgh
  • Chemical potential μ = G/n increases by Mgh: μ(h) = μ(0) + Mgh
  • Fugacity remains constant during this elevation process: f(h) = f(0)

Key insight: The additional Mgh term in the chemical potential expression accounts for the gravitational potential energy when vertical extent is considerable.

40

Internal Pressure

7.2 Internal Pressure

🧭 Overview

🧠 One-sentence thesis

In an equilibrium system of a pure substance, the chemical potential must be equal across all phases and elevations, which determines how pressure, fugacity, and phase coexistence vary with conditions like gravity and surface curvature.

📌 Key points (3–5)

  • Chemical potential equality: At equilibrium, the chemical potential has the same value everywhere in the system—across all phases and all elevations in a gravitational field.
  • Gravitational effects: In a tall gas column, fugacity and pressure decrease exponentially with elevation to maintain constant chemical potential (barometric formula).
  • Surface tension effects: Inside a small liquid droplet, pressure is higher than in the surrounding vapor due to surface tension (Laplace equation).
  • Degrees of freedom: A pure-substance equilibrium system always has exactly three independent variables total; the number of intensive independent variables (variance F) equals 3 − P, where P is the number of phases.
  • Common confusion: "Chemical potential" vs "total chemical potential"—this book defines chemical potential as molar Gibbs energy at a given elevation (constant at equilibrium), while some texts use "total chemical potential" for the elevation-dependent expression.

🌍 Chemical potential in gravitational fields

🧲 Chemical potential with elevation

Chemical potential of a pure substance at any elevation: the molar Gibbs energy at that elevation.

  • At equilibrium, the chemical potential μ(h) has the same value at each elevation h.
  • The expression for chemical potential at elevation h is:
    • μ(h) = μ°(g) + RT ln(f(h)/p°) + Mgh
    • Where M is molar mass, g is gravitational acceleration, h is elevation, f is fugacity, and p° is standard pressure.
  • The term Mgh accounts for gravitational potential energy; it is needed when the vertical extent of the gas is considerable.
  • When gravitational effects are negligible, the simpler form μ = μ°(g) + RT ln(f/p°) applies.

📉 How fugacity changes with elevation

  • For chemical potential to remain constant at all elevations, fugacity must decrease with increasing elevation.
  • By equating μ(h) at arbitrary elevation h with μ(0) at reference elevation zero:
    • f(h) = f(0) exp(−Mgh/RT)
  • This shows fugacity decreases exponentially as elevation increases.
  • Don't confuse: the chemical potential stays constant, but fugacity (and pressure) must vary to achieve this constancy.

🌡️ The barometric formula

  • If the gas behaves ideally (fugacity equals pressure), the fugacity relation becomes:
    • p(h) = p(0) exp(−Mgh/RT)
  • This is the barometric formula for a pure ideal gas.
  • It shows that in an equilibrium tall column of ideal gas, pressure decreases exponentially with increasing elevation.
  • Example: At sea level pressure p(0), the pressure at height h is lower by the exponential factor.

⚖️ Alternative derivation using forces

  • A shorter derivation uses Newton's second law instead of chemical potential.
  • Consider a thin horizontal slab of gas with density ρ, area As, and thickness Δh.
  • Three vertical forces act on the slab:
    • Upward pressure force pAs at the bottom
    • Downward pressure force (p + Δp)As at the top
    • Downward gravitational force ρgAs Δh
  • At equilibrium (net force zero): Δp = −ρg Δh
  • In the limit of infinitesimal thickness: dp = −ρg dh
  • This general relation applies to any fluid at equilibrium in a gravitational field.
  • For an ideal gas, substitute ρ = Mp/RT and integrate to recover the barometric formula.

💧 Pressure in liquid droplets

🔵 The Laplace equation

Laplace equation: pl = pg + 2γ/r, where pl is pressure inside the droplet, pg is gas pressure, γ is surface tension, and r is droplet radius.

  • A small spherical liquid droplet surrounded by its vapor has higher internal pressure than the surrounding vapor.
  • This pressure difference results from surface tension, which acts like a stretched membrane.
  • The surface tension minimizes the surface-to-volume ratio, creating a spherical shape.

📏 How droplet size affects pressure

  • The pressure difference is significant if r is small and decreases as r increases.
  • As r → ∞ (flat surface of bulk liquid), pl approaches pg—the pressure difference vanishes.
  • Example: A tiny droplet has much higher internal pressure than a large droplet of the same liquid at the same temperature.

🫧 Application to gas bubbles

  • The Laplace equation also applies to a gas bubble surrounded by liquid.
  • In this case, liquid and gas phases switch roles: pg = pl + 2γ/r
  • The gas inside the bubble has higher pressure than the surrounding liquid.

🧪 Derivation approach

  • The derivation considers a closed system with a spherical droplet and surrounding vapor, treated as open subsystems.
  • The infinitesimal change in internal energy includes contributions from:
    • Liquid phase: Tl dSl − pl dVl + μl dnl
    • Gas phase: Tg dSg − pg dVg + μg dng
    • Surface work: γ dAs (where As is surface area)
  • At equilibrium, liquid and gas have equal temperatures and equal chemical potentials.
  • The pressure difference depends on the radius r through the surface area term.
  • The Laplace equation is valid even when the liquid and gas are mixtures (not just pure substances).

🔢 Counting independent variables

🧮 Variables without equilibrium

  • In the absence of any equilibrium, each phase α has three independent variables: Tα, pα, and nα.
  • A system of P phases without thermal, mechanical, or transfer equilibrium would have 3P independent variables.
  • Note: P refers to the number of different kinds of phases, not the number of separate pieces (splitting a crystal doesn't change P if intensive properties remain the same).

🔗 Equilibrium constraints

  • Each independent equilibrium relation imposes a restriction and reduces the number of independent variables by one.
  • Thermal equilibrium: For P phases, there are P − 1 independent temperature relations (e.g., T′ = T″, T′ = T‴, etc.).
  • Mechanical equilibrium: P − 1 independent pressure relations.
  • Transfer equilibrium: P − 1 independent chemical potential relations.
  • Total number of independent equilibrium relations: 3(P − 1).

🎯 Result for pure substances

  • Number of independent variables at equilibrium = 3P − 3(P − 1) = 3.
  • An open single-substance system with any number of phases has exactly three independent variables at equilibrium.
  • Example: In a two-phase system at equilibrium, we may vary T, n′, and n″ independently; then p is determined (for given T, p must allow both phases to have the same chemical potential).

🎲 The Gibbs phase rule

📐 Degrees of freedom (variance)

Variance F (or number of degrees of freedom): the number of intensive independent variables in an equilibrium system.

  • A complete state description requires extensive variables (e.g., volume, mass, or amount of each phase) to specify how much of each phase is present.
  • For a pure substance with P phases and 3 total independent variables, the remaining 3 − P variables may be intensive.
  • The Gibbs phase rule for a pure substance: F = 3 − P
  • This is a special case (C = 1) of the general Gibbs phase rule F = C − P + 2 for multicomponent systems.

🔍 Two interpretations of variance

  1. Descriptive: F is the number of intensive variables needed to describe an equilibrium state, in addition to the amount of each phase.
  2. Operational: F is the maximum number of intensive properties we may vary independently while the phases remain in equilibrium.

🏷️ Terminology by number of phases

Number of phases PVariance F = 3 − PTermMeaning
12BivariantTwo intensive properties can vary independently
21UnivariantOnly one intensive property can vary independently
30InvariantNo intensive properties can vary independently
  • Example: A system of liquid and gaseous H₂O (two phases) is univariant (F = 3 − 2 = 1); we can independently vary only one intensive property, such as T, while liquid and gas remain in equilibrium—pressure is then fixed by the equilibrium condition.
41

Phase Diagrams of Pure Substances

7.3 Thermal Properties

🧭 Overview

🧠 One-sentence thesis

Phase diagrams map the equilibrium states of a pure substance across temperature, pressure, and volume, showing which phases exist under given conditions and how they coexist.

📌 Key points (3–5)

  • What phase diagrams show: the number and kinds of phases present at equilibrium for given values of independent variables (pressure, temperature, molar volume).
  • Single-phase vs multi-phase regions: a system point in a one-phase area means only that phase exists; a point in a two-phase area or on a coexistence curve means those phases coexist in equilibrium.
  • The triple point/line: the unique temperature and pressure where three phases (solid, liquid, gas) can coexist; it appears as a point on pressure–temperature diagrams and a line on pressure–volume diagrams.
  • Common confusion: the critical point is not just "high temperature"—it is the specific temperature above which no liquid–gas phase transition occurs, only a single supercritical fluid exists.
  • How to find phase properties: use tie lines on pressure–volume diagrams to read off molar volumes of coexisting phases and apply the lever rule to determine amounts in each phase.

📐 Understanding phase diagram structure

📐 Three-dimensional surface and projections

  • The full phase behavior of a pure substance is represented by a three-dimensional surface with axes for pressure (p), molar volume (V/n), and temperature (T).
  • Two-dimensional projections simplify analysis:
    • Pressure–volume diagram: projects the surface onto the p–(V/n) plane.
    • Pressure–temperature diagram: projects the surface onto the p–T plane.
  • Each projection highlights different information: pressure–volume shows molar volumes of phases; pressure–temperature shows coexistence curves and the triple point.

🗺️ Reading the system point

System point: the intersection of coordinates corresponding to given values of two independent variables on a phase diagram.

  • Place the system point at the coordinates of your chosen variables (e.g., a specific p and T).
  • The location tells you:
    • In a one-phase area: only that phase exists at equilibrium.
    • In a two-phase area or on a coexistence curve: those two phases coexist.
    • On the triple line (pressure–volume) or triple point (pressure–temperature): up to three phases can coexist.

🔢 Degrees of freedom (variance)

  • Single phase (bivariant): F = 3 − 1 = 2; you can independently vary two intensive properties (e.g., p and T) and stay in the one-phase region.
  • Two phases (univariant): F = 3 − 2 = 1; only one intensive variable (T or p) can be varied independently while both phases coexist.
  • Three phases (invariant): F = 3 − 3 = 0; there is only one unique temperature (triple-point temperature T_tp) and one pressure (triple-point pressure p_tp) for three-phase coexistence.

🔗 Two-phase equilibrium and coexistence

🔗 Coexistence curves and phase boundaries

Coexistence curve (phase boundary): a curve on the pressure–temperature diagram representing the locus of states where two phases coexist in equilibrium.

  • Because both phases share the same temperature and pressure (assuming no internal constraints), the two-phase area on a pressure–volume diagram projects onto a single curve on the pressure–temperature diagram.
  • Examples:
    • Vapor-pressure curve: shows how the vapor pressure of a liquid varies with temperature.
    • Boiling-point curve: shows how the boiling point varies with pressure (same curve, different perspective).
    • Melting-point curve: solid–liquid coexistence.

🌡️ Key temperatures and pressures

TermDefinition
Vapor pressure (saturation vapor pressure)Pressure at which solid–gas or liquid–gas phases are in equilibrium at a given temperature.
Sublimation pressureVapor pressure of a solid.
Melting point (freezing point)Temperature at which solid and liquid coexist at a given pressure.
Boiling point (saturation temperature)Temperature at which liquid and gas coexist at a given pressure.
Normal melting/boiling pointMelting or boiling point at 1 atm pressure.
Standard melting/boiling pointMelting or boiling point at 1 bar (standard pressure).
  • Example: Water's normal boiling point is 99.97 °C (at 1 atm, the "steam point"); its standard boiling point is 99.61 °C (at 1 bar, slightly lower pressure).

🔬 Measuring vapor pressure: the isoteniscope

  • A simple device to measure liquid vapor pressure at fixed temperature.
  • Procedure:
    1. Place liquid in the vessel and U-tube.
    2. Reduce pressure until liquid boils gently, sweeping out air with vapor.
    3. Adjust pressure until liquid levels in both limbs of the U-tube are equal.
    4. At this point, vapor pressure equals the measured pressure in the side tube.

🔺 The triple point and triple line

🔺 What the triple point represents

Triple point: the unique temperature and pressure at which three phases (solid, liquid, gas) can coexist in equilibrium for a pure substance.

  • On a pressure–temperature diagram, the triple line projects to a single point.
  • On a pressure–volume diagram, the triple line is a horizontal line segment showing the range of V/n values for the three coexisting phases.
  • The triple point is invariant (F = 0): T_tp and p_tp are fixed and unique to each substance.

🔺 Molar volumes at the triple line

  • The triple line on the pressure–volume diagram touches three one-phase areas:
    • The two ends correspond to the molar volumes of the solid and gas.
    • An intermediate position corresponds to the molar volume of the liquid.
  • If the system point is at either end, only one phase (of that molar volume) exists.
  • If the system point is anywhere between the ends, two or three phases may be present.

🔺 Multiple triple points at high pressure

  • Some substances (e.g., H₂O) have additional triple points at high pressures involving different solid phases (different crystal structures).
  • Example: Water has seven solid phases (ice I, II, III, V, VI, VII, VIII) and multiple triple points for combinations like two solids + liquid or three solids.
  • Ice I is ordinary ice, stable below ~2 bar; at higher pressures and temperatures above 273 K, "hot ice" (e.g., ice VI or VII) can exist.

Don't confuse: The triple point is not the same as the critical point; the triple point involves three phases at low-to-moderate conditions, while the critical point is the endpoint of the liquid–gas coexistence curve at high temperature.

🌀 The critical point and supercritical fluids

🌀 Defining the critical point

Critical temperature: the temperature above which only one fluid phase can exist, regardless of pressure or volume.

Critical point: the point on a phase diagram where liquid and gas coexist at the critical temperature; the pressure at this point is the critical pressure.

  • Above the critical temperature, no liquid–gas phase transition occurs.
  • The critical point is the endpoint of the liquid–gas coexistence curve on the pressure–temperature diagram.

🌀 Observing the critical point experimentally

  • Procedure:
    1. Evacuate a sealed glass vessel and introduce substance so V/n ≈ critical molar volume.
    2. Heat above the critical temperature → single fluid phase.
    3. Cool slowly to just above critical temperature → critical opalescence (cloudy appearance from light scattering due to large local density fluctuations).
    4. At the critical temperature, a meniscus forms between liquid and gas of nearly the same density.
    5. Further cooling → liquid density increases, gas density decreases.

🌀 Supercritical fluids

Supercritical fluid: a fluid at temperature and pressure both above the critical values; it does not undergo a phase transition to a different fluid phase when temperature or pressure is changed (at constant pressure or temperature, respectively).

  • Properties:
    • Density comparable to liquids.
    • More compressible than liquids.
    • Excellent solvent for solids and liquids; solvating power can be tuned by varying pressure or temperature.
  • Applications: chromatography, solvent extraction (easy removal as gas by reducing pressure).

Don't confuse: A supercritical fluid can still solidify if pressure is increased at constant temperature (e.g., supercritical water can become ice VII under isothermal compression).

🌀 Measuring critical properties

  • Critical temperature: observe appearance/disappearance of liquid–gas meniscus.
  • Critical pressure: measure pressure at critical temperature with a manometer.
  • Critical density: extrapolate the mean density of coexisting liquid and gas phases, (ρ_l + ρ_g)/2, to the critical temperature.

🌀 Law of rectilinear diameters

  • The mean density of coexisting liquid and gas phases is approximately a linear function of temperature.
  • Also called the law of Cailletet and Matthias.
  • Limitation: small deviations occur very close to the critical point (e.g., SF₆ data show slight nonlinearity), as predicted by recent theory.

🎚️ The lever rule: determining phase amounts

🎚️ What the lever rule calculates

  • When a system point lies in a two-phase area of a pressure–volume diagram, the lever rule determines the relative amounts (n_l and n_g) in each phase.
  • It applies to a fixed total amount n of substance divided into two phases at constant T and p.

🎚️ Using tie lines

Tie line: a horizontal line of constant pressure drawn through the system point in a two-phase area, extending to the boundaries with one-phase areas.

  • The ends of the tie line (points A and B) give the molar volumes of the two phases:
    • Left end (A): molar volume of liquid, V_l_m.
    • Right end (B): molar volume of gas, V_g_m.
  • The system point (S) lies somewhere on the tie line, with V/n between V_l_m and V_g_m.

🎚️ Lever rule formula

  • Define lengths (in units of V/n):

    • L_l = (V/n) − V_l_m (distance from left end A to system point S).
    • L_g = V_g_m − (V/n) (distance from system point S to right end B).
  • The lever rule states:

    n_l × L_l = n_g × L_g

    or equivalently:

    n_g / n_l = L_l / L_g

  • Interpretation: the ratio of amounts in the two phases is inversely proportional to the distances from the system point to the respective phase boundaries.

🎚️ Example application

  • Suppose the system point S is two-thirds of the way from the left end of the tie line.
  • Then L_l (distance to left) is twice L_g (distance to right): L_l = 2 × L_g.
  • Lever rule: n_g / n_l = L_l / L_g = 2.
  • Result: one-third of the total amount is liquid, two-thirds is gas.

Don't confuse: The lever rule cannot be applied to a point on the triple line (three phases), because knowing V/n alone is insufficient to determine the amounts in three phases.

🎚️ Physical interpretation

  • As heat is added at constant T and p, liquid vaporizes → system point moves right along the tie line → V increases.
  • As heat is removed, gas condenses → system point moves left → V decreases.
  • Molar volumes and other intensive properties of the individual phases remain constant during these changes.
42

Phase Diagrams, Phase Transitions, and Coexistence Curves

7.4 Heating at Constant Volume or Pressure

🧭 Overview

🧠 One-sentence thesis

The lever rule allows us to calculate the relative amounts of two coexisting phases from their positions on a phase diagram, while molar transition quantities describe the energy and entropy changes when a pure substance moves between phases at equilibrium.

📌 Key points (3–5)

  • Lever rule: For two coexisting phases on a tie line, the ratio of amounts in each phase equals the inverse ratio of distances from the system point to the ends of the tie line.
  • Molar transition quantities: These are changes in extensive properties (enthalpy, entropy, etc.) per amount transferred between phases at constant T and p; they depend on only one intensive variable because the system is univariant.
  • Standard vs. equilibrium transitions: Δ_vap H (without °) refers to real coexisting phases; Δ_vap H° (with °) involves a hypothetical ideal gas state and requires a multi-step calculation path.
  • Common confusion: Molar Gibbs energy of any equilibrium phase transition is always zero (Δ_trs G = 0), not the enthalpy or entropy changes.
  • Volume properties from diagrams: Isothermal compressibility and cubic expansion coefficient can be read from the slopes of isotherms and isobars on p–V phase diagrams.

📐 The Lever Rule

📐 What the lever rule tells us

Lever rule: For coexisting liquid and gas phases of a pure substance, n_l × L_l = n_g × L_g, or n_g / n_l = L_l / L_g.

  • What it calculates: the relative amounts (n_l and n_g) of two phases present when the system point lies on a tie line.
  • How it works: The system point S on a tie line divides the line into two segments; L_l is the distance from S to the gas end, L_g is the distance from S to the liquid end.
  • The ratio of amounts is the inverse ratio of distances.

🔍 How to apply it

  • The total volume and total amount must satisfy:
    • V = V_l + V_g = n_l V_l_m + n_g V_g_m
    • n = n_l + n_g
  • The value V/n at the system point is given by: V/n = (n_l V_l_m + n_g V_g_m) / (n_l + n_g).
  • Rearranging gives: n_l (V_l_m − V/n) = n_g (V/n − V_g_m).
  • The quantities V_l_m − V/n and V/n − V_g_m are the lengths L_l and L_g measured in units of V/n.

Example: If the system point S is two-thirds of the way from the left (liquid) end, then L_l is twice as long as L_g. The lever rule gives n_g / n_l = 2, so one-third of the total amount is liquid and two-thirds is gas.

⚠️ Limitations

  • Cannot apply to three phases: The lever rule does not work for a point on the triple line, because V/n alone is insufficient to determine the relative amounts in three phases.
  • General form: The lever rule can be generalized to any two-phase area of a two-dimensional phase diagram where tie lines are valid, using any two extensive state functions a and b, with F = a/b defining the system point position.

🌡️ Volume Properties from Phase Diagrams

🌡️ Reading isotherms (constant T curves)

  • Isothermal compressibility: κ_T = (−1/V_m) × (slope of isotherm).
  • The excerpt shows that isotherm slopes are large and negative in the liquid region, smaller and negative in the gas and supercritical fluid regions, and approach zero at the critical point.
  • Implication: The gas and supercritical fluid have much greater isothermal compressibility than the liquid; κ_T approaches infinity at the critical point.
  • Critical opalescence: Large density fluctuations at the critical point (where κ_T is large) cause the visible opalescence seen in experiments.

🌡️ Reading isobars (constant p curves)

  • Cubic expansion coefficient: α = (1/V_m) × (slope of isobar).
  • Isobar slopes are large and positive in the liquid region, smaller and positive in the gas and supercritical fluid regions, and approach zero at the critical point.
  • Implication: The gas and supercritical fluid have much larger cubic expansion coefficients than the liquid; α approaches infinity at the critical point.
  • The excerpt notes that the infinite α at the critical point means density distribution is greatly affected by temperature gradients in the critical region.

📊 Two-phase area

  • The area containing horizontal isotherm segments is the two-phase area for coexisting liquid and gas.
  • The boundary of this area is defined by a curve drawn through the ends of the horizontal segments.
  • One-phase liquid lies to the left, one-phase gas to the right, and the critical point at the top.

🔄 Molar Transition Quantities

🔄 What they represent

Molar transition quantity: the change of an extensive property divided by the amount transferred between phases at constant T and p.

  • Molar enthalpy of vaporization (Δ_vap H): enthalpy change per amount when liquid changes to gas at equilibrium conditions.
  • Also called "molar heat of vaporization" because at constant p, enthalpy change equals heat transferred.
  • Other transition quantities: Δ_sub H (sublimation, solid → gas), Δ_fus H (fusion, solid → liquid), Δ_trs H (any transition).
  • Molar enthalpies of vaporization, sublimation, and fusion are positive; reverse processes (condensation, freezing) have negative enthalpy changes.

🧮 Calculation from phase properties

For vaporization:

  • Δ_vap H = ΔH / n = H_g_m − H_l_m
  • The molar enthalpy of vaporization is the difference between the molar enthalpies of the two phases.
  • Key point: Although H_g_m and H_l_m each depend on two intensive variables (T and p), Δ_vap H depends on only one intensive variable, because the two phases are in transfer equilibrium and the system is univariant.
  • We may treat Δ_vap H as a function of T only.

🔄 Symbols for common transitions

SymbolMeaningDirection
Δ_vap HMolar enthalpy of vaporizationliquid → gas
Δ_sub HMolar enthalpy of sublimationsolid → gas
Δ_fus HMolar enthalpy of fusionsolid → liquid
Δ_trs HMolar enthalpy of any transitiongeneral

The same subscripts apply to other molar transition quantities (entropy, internal energy, etc.).

⚖️ Special Case: Molar Gibbs Energy of Transition

⚖️ Why it is always zero

For a phase transition ' → ":

  • Δ_trs G = G"_m − G'_m = μ" − μ'
  • But the transition is between two phases at equilibrium, requiring both phases to have the same chemical potential: μ" − μ' = 0.
  • Therefore: Δ_trs G = 0 for any equilibrium phase transition of a pure substance.

🔗 Relationship between enthalpy and entropy

Since G = H − TS, we can write:

  • Δ_trs G = Δ_trs H − T_trs Δ_trs S
  • Setting Δ_trs G = 0 gives:

Δ_trs S = Δ_trs H / T_trs

  • This relates the molar entropy and molar enthalpy of an equilibrium phase transition at the transition temperature T_trs.
  • Alternative derivation: For a reversible transition, the second law gives ΔS = q / T_trs = ΔH / T_trs; dividing by the amount transferred gives the same result.

⚠️ Don't confuse

  • Δ_trs G = 0 does not mean Δ_trs H = 0 or Δ_trs S = 0.
  • The enthalpy and entropy changes are non-zero but related by the transition temperature.

🔬 Measuring Transition Enthalpies

🔬 Calorimetric method with electrical work

  • Setup: A constant-pressure adiabatic calorimeter containing coexisting phases.
  • Procedure: Perform a known quantity of electrical work (I² R_el Δt) on the system; measure the resulting amount of substance transferred between phases.
  • Why it works: The first law shows that the electrical work equals the heat that would be needed to cause the same change of state; at constant p, this heat is the enthalpy change.
  • Advantage: This method is similar to measuring heat capacity at constant pressure, but simpler because temperature remains constant and no correction for calorimeter heat capacity is needed.

🎯 Standard Molar Transition Quantities

🎯 Definition and the hypothetical state problem

Standard molar enthalpy of vaporization (Δ_vap H°): enthalpy change when pure liquid in its standard state at a specified temperature changes to gas in its standard state at the same temperature, divided by the amount changed.

  • Initial state: real (pure liquid at pressure p°).
  • Final state: hypothetical (gas behaving ideally at pressure p°).
  • Problem: The liquid and gas are not necessarily in equilibrium with one another at pressure p° and the given temperature.
  • We cannot evaluate Δ_vap H° directly from a calorimetric measurement without further corrections.
  • The same difficulty applies to Δ_sub H°.

🔄 Contrast with non-standard quantities

  • Δ_vap H and Δ_sub H (without °) refer to reversible transitions between two real phases coexisting in equilibrium.
  • Δ_fus H° also refers to real phases because both solid and liquid can coexist at p°.

🛤️ Three-step calculation path

To evaluate Δ_vap X° or Δ_sub X° at a given temperature T:

  1. Isothermal pressure change of condensed phase: Start with the standard state at p°; end at the vapor pressure p_vap at temperature T. Calculate ΔX_m from formulas or approximations in tables.
  2. Reversible vaporization or sublimation: Form the real gas at T and p_vap. The change is Δ_vap X or Δ_sub X, which can be evaluated experimentally.
  3. Isothermal change of real gas to ideal gas: Change from real gas at p_vap to hypothetical ideal gas at p°. Use formulas relating molar quantities of a real gas to standard molar quantities.

The sum of ΔX_m for these three steps is Δ_vap X° or Δ_sub X°.

🌐 Chemical Potential Surfaces and Coexistence

🌐 Visualizing equilibrium

  • The chemical potential μ of a pure substance in a single phase can be treated as a function of T and p, represented by a three-dimensional surface.
  • Equilibrium condition: Two phases can coexist in equilibrium only when they have the same T, p, and μ.
  • Geometric interpretation: Equilibrium in a two-phase system exists only along the intersection of the chemical potential surfaces of the two phases.
  • The shape of each surface is determined by the partial derivatives of μ with respect to T and p.

🗺️ Coexistence curves

  • A coexistence curve on a pressure–temperature phase diagram shows the conditions (T and p values) under which two phases can coexist in equilibrium.
  • These curves correspond to the projection of the intersection line of two chemical potential surfaces onto the T–p plane.
43

Partial Derivatives with Respect to T, p, and V

7.5 Partial Derivatives with Respect to T, p, and V

🧭 Overview

🧠 One-sentence thesis

The Clapeyron and Clausius–Clapeyron equations relate the slope of phase coexistence curves to measurable thermodynamic quantities, enabling the calculation of transition enthalpies and the prediction of how phase equilibria shift with temperature and pressure.

📌 Key points (3–5)

  • What coexistence curves represent: conditions under which two phases can coexist in equilibrium, determined by the intersection of chemical potential surfaces where both phases have the same T, p, and μ.
  • The Clapeyron equation: an exact relation giving the slope of any coexistence curve as dp/dT = ΔtrsH/(T ΔtrsV), valid for all phase transitions without approximation.
  • The Clausius–Clapeyron equation: an approximate form for liquid–gas or solid–gas equilibria that assumes the gas volume dominates and behaves ideally, yielding d ln(p/p°)/dT ≈ ΔtrsH/(RT²).
  • Common confusion: Clapeyron vs. Clausius–Clapeyron—the former is exact and applies to all transitions (including solid–liquid), while the latter is approximate and only valid for vaporization/sublimation away from the critical point.
  • Why it matters: these equations allow experimental determination of transition enthalpies from vapor pressure measurements and enable prediction of melting/boiling points at different pressures.

🔬 Calorimetric measurement of transition enthalpies

🔬 Electrical work method

The most precise measurement of the molar enthalpy of an equilibrium phase transition uses electrical work performed on a system containing coexisting phases in a constant-pressure adiabatic calorimeter.

  • A known quantity of electrical work (I² Rₑₗ Δt) is performed on the system.
  • The resulting amount of substance transferred between phases is measured.
  • The first law shows that this electrical work equals the heat that would be needed to cause the same change of state.
  • At constant pressure, this heat equals the enthalpy change of the process.
  • Example: heating a liquid–vapor mixture electrically causes a measurable amount of liquid to vaporize; the work divided by the amount vaporized gives the molar enthalpy of vaporization.
  • The method is similar to measuring heat capacity (Sec. 7.3.2), but here temperature remains constant and no calorimeter heat capacity correction is needed.

📐 Standard molar transition quantities

📐 Definition and hypothetical states

Standard molar enthalpy of vaporization (ΔvapH°) is the enthalpy change when pure liquid in its standard state changes to gas in its standard state at the same temperature, divided by the amount changed.

  • The initial state (pure liquid at pressure p°) is real.
  • The final state (gas behaving ideally at pressure p°) is hypothetical.
  • The liquid and gas are not necessarily in equilibrium at p° and the temperature of interest.
  • Cannot evaluate ΔvapH° directly from calorimetric measurement without further corrections.
  • The same difficulty applies to ΔsubH° (standard sublimation enthalpy).

📐 Real vs. standard transitions

Don't confuse: ΔvapH (without °) refers to a reversible transition between two real phases coexisting in equilibrium, while ΔvapH° involves a hypothetical ideal gas state.

QuantityInitial stateFinal statePhases in equilibrium?
ΔvapH, ΔsubH, ΔfusH°Real phase at pvapReal phase at pvapYes
ΔvapH°, ΔsubH°Real at p°Hypothetical ideal at p°No

📐 Three-step path for evaluation

Standard molar transition quantities (ΔvapX° or ΔsubX°) are functions only of T. To evaluate them at a given temperature, calculate the change of Xₘ for a path connecting the standard state of the condensed phase with that of the gas:

  1. Isothermal pressure change of the liquid or solid from p° to pvap (vapor pressure at T).
    • Use expressions from Table 7.4 to find ΔXₘ.
  2. Reversible vaporization or sublimation to form real gas at T and pvap.
    • ΔXₘ = ΔvapX or ΔsubX (experimentally measurable).
  3. Isothermal change of real gas at pvap to hypothetical ideal gas at p°.
    • Use formulas from Table 7.5 relating real gas to standard quantities.

The sum of ΔXₘ for these three steps gives ΔvapX° or ΔsubX°.

🗺️ Chemical potential surfaces and coexistence

🗺️ Three-dimensional representation

The chemical potential μ of a pure substance in a single phase can be treated as a function of T and p, represented by a three-dimensional surface.

  • Equilibrium between two phases requires both phases have the same T, p, and μ.
  • Two-phase equilibrium exists only along the intersection of the surfaces of the two phases.
  • The shape of each surface is determined by partial derivatives: (∂μ/∂T)ₚ = −Sₘ and (∂μ/∂p)ₜ = Vₘ.

🗺️ Temperature dependence at constant pressure

At constant p, the slope of μ versus T is negative because molar entropy is always positive: (∂μ/∂T)ₚ = −Sₘ.

  • The magnitude of the slope increases on going from solid to liquid to gas.
  • This is because molar entropies of sublimation and vaporization are positive (Sₘ(gas) > Sₘ(liquid) > Sₘ(solid)).
  • The stable phase at each temperature is the one of lowest μ (transfer from higher to lower μ at constant T and p is spontaneous).
  • Example: For H₂O at 0.03 bar (above triple-point pressure), solid and liquid curves intersect at a melting point (A) and liquid and gas curves intersect at a boiling point (B).

🗺️ Pressure dependence at constant temperature

From (∂μ/∂p)ₜ = Vₘ, a pressure reduction at constant temperature lowers the chemical potential of a phase.

  • The shift is proportional to the molar volume of the phase.
  • For condensed phases (solid, liquid), the shift is small because Vₘ is small.
  • For the gas phase, the shift is substantial because Vₘ(gas) is large.
  • Example: Reducing pressure from 0.03 bar to 0.003 bar (below H₂O triple-point pressure) shifts the gas curve downward enough to intersect the solid curve at a sublimation point (C), while solid and liquid curves shift negligibly.

🗺️ Below the triple-point pressure

At any pressure below the triple-point pressure, only solid–gas equilibrium is possible.

  • The liquid phase is not stable at any temperature below the triple-point pressure.
  • This is shown in the pressure–temperature phase diagram.

⚖️ The Clapeyron equation

⚖️ Derivation from equilibrium condition

If two coexisting phases α and β are changed in temperature by dT without changing pressure, they will no longer be in equilibrium because their chemical potentials change unequally.

  • For phases to remain in equilibrium during temperature change dT, there must be a simultaneous pressure change dp.
  • The changes must satisfy: dμᵅ = dμᵝ.
  • Using dμ = −Sₘ dT + Vₘ dp (Eq. 7.8.2): −Sₘᵅ dT + Vₘᵅ dp = −Sₘᵝ dT + Vₘᵝ dp.

Rearranging gives:

Clapeyron equation: dp/dT = ΔtrsS/ΔtrsV = ΔtrsH/(T ΔtrsV)

where ΔtrsS = Sₘᵝ − Sₘᵅ and ΔtrsV = Vₘᵝ − Vₘᵅ.

  • This equation contains no approximations.
  • Valid for any phase transition of a pure substance.
  • The alternative form uses ΔtrsS = ΔtrsH/Ttrs (Eq. 8.3.5).

⚖️ Sign of the slope

The Clapeyron equation gives the slope of the coexistence curve as a function of measurable quantities.

TransitionΔtrsHΔtrsVSlope sign
Sublimation (s→g)PositivePositivePositive
Vaporization (l→g)PositivePositivePositive
Fusion (s→l)PositivePositive or negativePositive or negative
  • For sublimation and vaporization, both ΔtrsH and ΔtrsV are positive → positive slope.
  • For fusion, ΔfusH is always positive, but ΔfusV may be positive (most substances expand on melting) or negative (e.g., H₂O, Bi, Sb contract on melting).
  • The absolute value of ΔfusV is small → solid–liquid coexistence curve is relatively steep.

⚖️ Integration for solid–liquid equilibria

Because the cubic expansion coefficient and isothermal compressibility of condensed phases are relatively small, ΔfusV is approximately constant for small changes of T and p.

If ΔfusH is also practically constant, integrating dp = (ΔfusH/ΔfusV)(dT/T) yields:

  • p₂ − p₁ ≈ (ΔfusH/ΔfusV) ln(T₂/T₁)
  • Or: T₂ ≈ T₁ exp[ΔfusV(p₂ − p₁)/ΔfusH]

This allows estimation of how the melting point depends on pressure.

Don't confuse: This integration is only valid for solid–liquid equilibria where volume changes are small; it does not apply to liquid–gas or solid–gas equilibria.

🌡️ The Clausius–Clapeyron equation

🌡️ Approximations for gas–condensed phase equilibria

When a gas phase coexists with a liquid or solid phase, and T and p are not close to the critical point, the molar volume of the gas is much greater than that of the condensed phase.

  • For vaporization: ΔvapV = Vₘ(g) − Vₘ(l) ≈ Vₘ(g)
  • For sublimation: ΔsubV = Vₘ(g) − Vₘ(s) ≈ Vₘ(g)
  • Further approximation: the gas behaves as an ideal gas, so Vₘ(g) ≈ RT/p.

Substituting into the Clapeyron equation gives:

Clausius–Clapeyron equation: dp/dT ≈ p ΔtrsH/(RT²)

  • Valid for vaporization or sublimation only.
  • Not valid for solid–liquid equilibria or for liquid–gas near the critical point.

🌡️ Alternative forms

The Clausius–Clapeyron equation can be written in several equivalent forms:

  • d ln(p/p°)/dT ≈ ΔtrsH/(RT²)
  • d ln(p/p°) ≈ (ΔtrsH/R) d(1/T)
  • d ln(p/p°)/d(1/T) ≈ −ΔtrsH/R

where p/p° is the dimensionless ratio of vapor pressure to standard pressure.

🌡️ Practical applications

Determining transition enthalpies from vapor pressure data:

A plot of ln(p/p°) versus 1/T has a slope at each temperature equal to −ΔvapH/R or −ΔsubH/R.

  • This provides an alternative to calorimetry for evaluating molar enthalpies of vaporization and sublimation.
  • For the recommended standard pressure of 1 bar, p/p° is simply the numerical value of p in bars.
  • Plots of ln(p/bar), ln(p/Pa), or ln(p/Torr) versus 1/T all have the same slope (but different intercepts) and yield the same ΔtrsH.

Estimating vapor pressures at different temperatures:

If ΔvapH or ΔsubH is essentially constant over a temperature range, integrating d ln(p/p°) = (ΔtrsH/R) d(1/T) gives:

ln(p₂/p₁) ≈ (ΔtrsH/R)[1/T₂ − 1/T₁]

  • Allows estimation of any one of p₁, p₂, T₁, T₂, or ΔtrsH given the other four.
  • Example: knowing the vapor pressure at one temperature and the enthalpy of vaporization, you can estimate the vapor pressure at another temperature.

🌡️ Triple-point relationships

At the triple point, all three equilibrium phase transitions (fusion, vaporization, sublimation) are possible.

  • Fusion followed by vaporization gives net sublimation: ΔfusH + ΔvapH = ΔsubH (at the triple point).
  • Since all three transition enthalpies are positive, ΔsubH > ΔvapH at the triple point.
  • According to the Clausius–Clapeyron equation, the slope of the solid–gas coexistence curve at the triple point is slightly greater than the slope of the liquid–gas coexistence curve.

Budget: 1000000 Tokens used: 19419 Tokens remaining: 980581 Token usage: 1.9%

44

7.6 Isothermal Pressure Changes

7.6 Isothermal Pressure Changes

🧭 Overview

🧠 One-sentence thesis

Calorimetric measurements using electrical work provide precise determination of phase transition enthalpies, while standard molar transition quantities require multi-step paths to connect real and hypothetical standard states.

📌 Key points (3–5)

  • Calorimetric method: electrical work performed on coexisting phases in an adiabatic calorimeter measures transition enthalpy directly.
  • Standard vs. non-standard transitions: standard molar quantities (with ı symbol) involve hypothetical ideal gas states, while non-standard quantities refer to real equilibrium phases.
  • Multi-step path requirement: evaluating standard molar transition quantities requires calculating changes through isothermal pressure adjustments and phase transitions.
  • Common confusion: standard molar vaporization enthalpy cannot be measured directly from calorimetry because the final state (ideal gas at standard pressure) is hypothetical, not real.
  • Coexistence curves: equilibrium between two phases occurs only where their chemical potential surfaces intersect at the same temperature and pressure.

🔬 Calorimetric measurement principles

⚡ Electrical work method

The most precise measurement of the molar enthalpy of an equilibrium phase transition uses electrical work performed on a system containing coexisting phases in a constant-pressure adiabatic calorimeter.

  • A known quantity of electrical work (I²R_el·Δt) is performed on the system.
  • The amount of substance transferred between coexisting phases is measured.
  • The first law shows that this electrical work equals the heat that would be needed to cause the same change of state.
  • At constant pressure, this heat equals the enthalpy change of the process.

🌡️ Constant temperature operation

  • The temperature remains constant during the measurement (unlike heat capacity measurements where temperature changes).
  • No correction for the heat capacity of the calorimeter is needed.
  • The method is similar to measuring heat capacity at constant pressure but adapted for phase transitions.
  • Example: electrical work causes liquid to vaporize while temperature stays fixed; measuring how much substance vaporizes gives the transition enthalpy.

🎯 Standard molar transition quantities

📐 Definition and characteristics

The standard molar enthalpy of vaporization, Δ_vap H°, is the enthalpy change when pure liquid in its standard state at a specified temperature changes to gas in its standard state at the same temperature, divided by the amount changed.

  • The initial state is real: pure liquid at standard pressure p°.
  • The final state is hypothetical: gas behaving ideally at standard pressure p°.
  • The liquid and gas are not necessarily in equilibrium with one another at pressure p° and the specified temperature.
  • Standard molar transition quantities (with ° symbol) are functions only of temperature T.

⚠️ Measurement limitation

  • Cannot evaluate directly from calorimetry: standard molar vaporization enthalpy (Δ_vap H°) and standard molar sublimation enthalpy (Δ_sub H°) require further corrections beyond simple calorimetric measurement.
  • Why: the calorimetric method measures transitions between real coexisting phases, not transitions ending in hypothetical ideal gas states.
  • Don't confuse: Δ_vap H and Δ_sub H (without ° symbol) refer to reversible transitions between two real phases coexisting in equilibrium and can be measured directly.

✅ Fusion exception

  • Standard molar fusion enthalpy (Δ_fus H°) refers to a reversible transition between two real phases coexisting in equilibrium.
  • This can be measured directly because both solid and liquid phases are real at the transition conditions.

🛤️ Multi-step path for standard quantities

🗺️ Three-step isothermal path

To evaluate standard molar transition quantities (Δ_vap X° or Δ_sub X°) at a given temperature, calculate the change of molar property X_m through these constant-temperature steps:

StepProcessHow to calculate ΔX_m
1Isothermal pressure change of liquid/solid from p° to vapor pressure p_vapUse expressions from Table 7.4 (exact or approximation)
2Reversible vaporization/sublimation to real gas at T and p_vapUse Δ_vap X or Δ_sub X from experimental measurement
3Isothermal change of real gas at p_vap to hypothetical ideal gas at p°Use formulas from Table 7.5 relating real gas to standard quantities

🔗 Path logic

  • The path connects the standard state of the condensed phase (liquid or solid at p°) with the standard state of the gas (ideal gas at p°).
  • The sum of ΔX_m for all three steps equals the desired standard molar transition quantity.
  • Example: to find standard molar vaporization entropy, start with liquid at standard pressure, adjust pressure to the vapor pressure, vaporize to real gas, then adjust the real gas to ideal gas behavior at standard pressure.

🧮 Applicable properties

  • X can represent any thermodynamic potential or the entropy of a phase.
  • Standard molar transition quantities include: Δ_vap X° = X°_m(g) - X_m(l) and Δ_sub X° = X°_m(g) - X_m(s).
  • These are functions only of temperature T.

📈 Coexistence curves and chemical potential

📊 Coexistence curve definition

A coexistence curve on a pressure–temperature phase diagram shows the conditions under which two phases can coexist in equilibrium.

  • These curves represent the specific combinations of temperature and pressure where two phases are stable together.
  • The conditions are explained by the equilibrium requirement that both phases have the same chemical potential.

🏔️ Chemical potential surfaces

  • The chemical potential μ of a pure substance in a single phase can be treated as a function of independent variables T and p.
  • This function is represented by a three-dimensional surface.
  • The shape of each phase's surface is determined by partial derivatives of chemical potential with respect to temperature and pressure.

⚖️ Equilibrium intersection

  • Equilibrium condition: both phases must have the same T, p, and μ.
  • Equilibrium in a two-phase system can exist only along the intersection of the surfaces of the two phases.
  • Don't confuse: equilibrium is not just matching temperature and pressure; the chemical potentials must also be equal, which happens only at specific T-p combinations forming the coexistence curve.
  • Example: the liquid and gas surfaces intersect along a curve; points on this curve represent the vapor pressure at each temperature where liquid and gas coexist.
45

Standard States of Pure Substances

7.7 Standard States of Pure Substances

🧭 Overview

🧠 One-sentence thesis

Standard molar transition quantities (like standard enthalpy of vaporization) require a multi-step path calculation because they connect real condensed phases to hypothetical ideal-gas states, unlike ordinary transition quantities that describe equilibrium between two real phases.

📌 Key points (3–5)

  • Calorimetric measurement: electrical work in an adiabatic calorimeter measures transition enthalpies by quantifying the energy needed to transfer substance between coexisting phases.
  • Standard vs. ordinary transition quantities: standard quantities (with ° symbol) involve hypothetical ideal-gas states and require path calculations; ordinary quantities describe real equilibrium transitions.
  • Three-step path: evaluating standard transition quantities requires (1) pressure change of condensed phase to vapor pressure, (2) reversible phase transition at equilibrium, (3) conversion of real gas to ideal gas at standard pressure.
  • Common confusion: standard molar enthalpy of vaporization cannot be measured directly from calorimetry because the final state (ideal gas at standard pressure) is hypothetical, not an equilibrium state with the liquid.
  • Chemical potential surfaces: coexistence curves arise from intersections of chemical potential surfaces, with slopes determined by molar entropy (temperature) and molar volume (pressure).

🔬 Measuring transition enthalpies

⚡ Electrical work method

Calorimetric measurement of transition enthalpies: a known quantity of electrical work is performed on a system containing coexisting phases in a constant-pressure adiabatic calorimeter, and the resulting amount of substance transferred between phases is measured.

  • The electrical work (I²R·t, where I is current, R is resistance, t is time) equals the heat needed to cause the same state change.
  • At constant pressure, this heat equals the enthalpy change of the process.
  • The method is similar to measuring heat capacity, but temperature remains constant during phase transition, so no calorimeter heat-capacity correction is needed.
  • Example: applying electrical work to ice-water mixture transfers a measurable amount of ice to water; the work divided by the amount gives the molar enthalpy of fusion.

🎯 Standard vs. ordinary transition quantities

🎯 What makes standard quantities different

The excerpt distinguishes two types of transition quantities:

TypeSymbol exampleInitial stateFinal stateMeasurable directly?
StandardΔ_vap H°Real liquid at p°Hypothetical ideal gas at p°No—requires path calculation
OrdinaryΔ_vap HReal liquid at p_vapReal gas at p_vapYes—from calorimetry at equilibrium
  • Standard molar enthalpy of vaporization (Δ_vap H°): enthalpy change when pure liquid in its standard state at specified temperature changes to gas in its standard state at the same temperature, divided by the amount changed.
  • The initial state (pure liquid at pressure p°) is real, but the final state (gas behaving ideally at pressure p°) is hypothetical.
  • The liquid and gas are not necessarily in equilibrium with one another at pressure p° and the temperature of interest.

🔄 Which transitions involve real states

  • Δ_vap H and Δ_sub H (without ° symbol): refer to reversible transitions between two real phases coexisting in equilibrium.
  • Δ_fus H° (fusion): also refers to equilibrium between two real phases.
  • Don't confuse: the ° symbol indicates involvement of a hypothetical ideal-gas state for vaporization and sublimation, but not for fusion (solid-liquid equilibrium is always between real phases).

🛤️ Three-step path for standard quantities

📐 Why a path is needed

Standard molar transition quantities (Δ_vap X° or Δ_sub X°, where X represents a thermodynamic potential or entropy) are functions only of temperature T.

To evaluate these at a given temperature, we must calculate the change of the molar quantity X_m for a path connecting the standard state of the liquid or solid with that of the gas.

🪜 The three steps at constant temperature

Step 1: Pressure change of condensed phase

  • Isothermal change of pressure of the liquid or solid, starting at standard pressure p° and ending at the vapor pressure p_vap at temperature T.
  • The change ΔX_m can be obtained from expressions relating molar quantities to pressure changes.

Step 2: Reversible phase transition

  • Reversible vaporization or sublimation to form the real gas at T and p_vap.
  • The change of X_m in this step is either Δ_vap X or Δ_sub X, which can be evaluated experimentally (from calorimetry).

Step 3: Real gas to ideal gas

  • Isothermal change of the real gas at pressure p_vap to the hypothetical ideal gas at pressure p°.
  • Formulas relate molar quantities of a real gas to the corresponding standard molar quantities.

Result: The sum of ΔX_m for these three steps is the desired standard quantity Δ_vap X° or Δ_sub X°.

Example: To find standard enthalpy of vaporization of water at 350 K, you cannot simply measure heat at standard pressure (water and steam may not be in equilibrium there); instead, calculate enthalpy changes for (1) compressing or expanding liquid to its vapor pressure at 350 K, (2) vaporizing at that equilibrium pressure, (3) converting real steam to ideal gas at standard pressure.

📈 Chemical potential and coexistence

🌐 Chemical potential surfaces

Chemical potential μ of a pure substance in a single phase: treated as a function of independent variables T and p, represented by a three-dimensional surface.

  • Equilibrium between two phases requires both phases have the same T, p, and μ.
  • Two-phase equilibrium exists only along the intersection of the chemical potential surfaces of the two phases.
  • A coexistence curve on a pressure-temperature phase diagram shows conditions under which two phases can coexist in equilibrium.

📉 How μ varies with temperature and pressure

The shape of each phase's surface is determined by partial derivatives:

  • (∂μ/∂T) at constant p = –S_m (negative molar entropy)
  • (∂μ/∂p) at constant T = V_m (molar volume)

Temperature dependence at constant pressure:

  • The slope of μ versus T is always negative (since molar entropy is always positive).
  • The magnitude of the slope increases from solid → liquid → gas, because molar entropies of sublimation and vaporization are positive.
  • The stable phase at each temperature is the one of lowest μ (transfer from higher to lower μ at constant T and p is spontaneous).

Pressure dependence at constant temperature:

  • A pressure reduction at constant temperature lowers the chemical potential of a phase.
  • The amount of lowering is proportional to the molar volume of the phase.

🔺 Triple point and phase intersections

Example (H₂O):

  • Triple-point pressure: 0.0062 bar.
  • At 0.03 bar (above triple-point pressure): solid and liquid curves intersect at a melting point (point A); liquid and gas curves intersect at a boiling point (point B).
  • At 0.003 bar (below triple-point pressure): pressure reduction shifts each curve downward by a distance proportional to the phase's molar volume; shifts of solid and liquid curves are much smaller than the gas curve shift (because molar volumes of condensed phases are much smaller).

Don't confuse: the intersection of two chemical potential surfaces in 3D (μ-T-p space) projects onto the p-T plane as a coexistence curve; this curve is not arbitrary but determined by the thermodynamic properties (S_m and V_m) of each phase.

46

Chemical Potential and Fugacity

7.8 Chemical Potential and Fugacity

🧭 Overview

🧠 One-sentence thesis

Chemical potential surfaces of different phases intersect along coexistence curves where two phases can equilibrate, and the slope of these curves is governed by how entropy and volume differences between phases respond to temperature and pressure changes.

📌 Key points

  • Chemical potential as equilibrium criterion: Two phases coexist in equilibrium only when they share the same temperature, pressure, and chemical potential.
  • How chemical potential varies: At constant pressure, chemical potential decreases with temperature (slope = –S_m); at constant temperature, it decreases with pressure (slope = V_m).
  • Phase stability rule: The stable phase at any condition is the one with the lowest chemical potential; transfer from higher to lower chemical potential is spontaneous.
  • Common confusion: Pressure changes shift all phase curves downward, but the gas curve shifts much more than solid or liquid curves because gas has much larger molar volume—this can change which phases intersect.
  • The Clapeyron equation: To keep two phases in equilibrium during a temperature change dT, pressure must change by dp such that both phases' chemical potentials change equally.

🌐 Chemical potential surfaces and phase equilibrium

🗺️ Three-dimensional representation

Chemical potential μ of a pure substance in a single phase: a function of independent variables T and p, represented as a three-dimensional surface.

  • Each phase (solid, liquid, gas) has its own μ(T, p) surface.
  • Equilibrium between two phases requires:
    • Same temperature T
    • Same pressure p
    • Same chemical potential μ
  • Geometric meaning: Two phases can coexist only along the intersection of their chemical potential surfaces.
  • Example: Figure 8.12 shows liquid and gas surfaces for H₂O intersecting along a curve; projecting this intersection onto the p–T plane generates the coexistence curve (the familiar phase boundary).

📐 Partial derivatives governing surface shape

The shape of each phase's chemical potential surface is determined by two fundamental relations:

DerivativeEqualsPhysical meaning
(∂μ/∂T) at constant p–S_mHow μ changes with temperature at fixed pressure
(∂μ/∂p) at constant TV_mHow μ changes with pressure at fixed temperature
  • Both derivatives come from the infinitesimal change: dμ = –S_m dT + V_m dp.
  • These relations (Eqs. 7.8.3 and 7.8.4) control how the surface tilts in the T and p directions.

🌡️ Temperature dependence of chemical potential

📉 Negative slope with temperature

  • At constant pressure, (∂μ/∂T)_p = –S_m.
  • Since molar entropy S_m is always positive, the slope of μ versus T is always negative.
  • Implication: Chemical potential decreases as temperature increases (at fixed pressure).

🔄 Different slopes for different phases

  • The magnitude of the slope increases going from solid → liquid → gas.
  • Reason: Molar entropies of sublimation and vaporization are positive, so S_m(gas) > S_m(liquid) > S_m(solid).
  • Example: In Figure 8.13(a) for H₂O at 0.03 bar:
    • The solid curve has the smallest (least negative) slope.
    • The liquid curve has a steeper slope.
    • The gas curve has the steepest slope.
  • Result: The curves intersect at specific temperatures:
    • Solid and liquid intersect at the melting point (point A).
    • Liquid and gas intersect at the boiling point (point B).

🏆 Phase stability and lowest chemical potential

  • At any given T and p, the stable phase is the one with the lowest μ.
  • Transfer of substance from higher μ to lower μ at constant T and p is spontaneous.
  • Don't confuse: A phase can have a well-defined chemical potential even when it's not the stable phase; stability is determined by comparison.

🔽 Pressure dependence of chemical potential

📊 How pressure shifts the curves

  • At constant temperature, (∂μ/∂p)_T = V_m.
  • A pressure reduction lowers the chemical potential of a phase.
  • The amount of downward shift is proportional to the molar volume V_m of the phase.

🎯 Unequal shifts and phase diagram changes

Example from H₂O (Figure 8.13):

  • At 0.03 bar (above triple-point pressure of 0.0062 bar):
    • Solid and liquid curves intersect at melting point A.
    • Liquid and gas curves intersect at boiling point B.
  • When pressure is reduced from 0.03 bar to 0.003 bar (below triple-point pressure):
    • All three curves shift downward.
    • Solid and liquid shifts are tiny (~0.002 kJ/mol) because their molar volumes are small.
    • Gas curve shifts substantially because gas has a large molar volume.
    • Result: The gas curve now intersects the solid curve at a sublimation point (point C), but no longer intersects the liquid curve.
  • Conclusion: At pressures below the triple-point pressure, only solid–gas equilibrium is possible; the liquid phase is not stable at any temperature.

⚠️ Common confusion: why gas shifts more

  • Don't confuse absolute molar volume with the shift magnitude: the shift Δμ is proportional to V_m × Δp.
  • Gas has V_m orders of magnitude larger than solid or liquid, so even the same pressure change causes a much larger shift in the gas curve.
  • This is why reducing pressure below the triple point eliminates the liquid phase entirely for H₂O.

📏 The Clapeyron equation

🔗 Maintaining equilibrium during changes

  • Start with two coexisting phases α and β at equilibrium (same T, p, μ).
  • If we change temperature by dT without changing pressure, the phases will no longer be in equilibrium because their chemical potentials change unequally (different S_m values).
  • Question: What simultaneous pressure change dp is needed so the phases remain in equilibrium?

⚖️ Equal chemical potential changes

For the phases to remain in equilibrium:

  • The change in chemical potential must be equal for both phases: dμ^α = dμ^β.
  • Using dμ = –S_m dT + V_m dp for each phase:
    • –S_m^α dT + V_m^α dp = –S_m^β dT + V_m^β dp
  • Rearranging:
    • (V_m^α – V_m^β) dp = (S_m^α – S_m^β) dT
  • This relation (Eq. 8.4.2) is the starting point for the Clapeyron equation.

🧮 Physical meaning

  • The required dp/dT ratio along the coexistence curve depends on the differences in molar entropy and molar volume between the two phases.
  • Example: For liquid–gas equilibrium, S_m(gas) >> S_m(liquid) and V_m(gas) >> V_m(liquid), so the slope dp/dT of the boiling curve is determined by these differences.
  • Don't confuse: This is not about how μ changes in a single phase, but about how the coexistence conditions (the intersection curve) shift when both T and p change together.
47

Standard Molar Quantities of a Gas

7.9 Standard Molar Quantities of a Gas

🧭 Overview

🧠 One-sentence thesis

Chemical potential changes with temperature and pressure determine which phase is stable and how coexistence curves behave, with the Clapeyron equation quantifying the slope of these curves using measurable transition properties.

📌 Key points (3–5)

  • Chemical potential and phase stability: at constant T and p, the phase with the lowest chemical potential is stable; transfer from higher to lower chemical potential is spontaneous.
  • How chemical potential varies: at constant p, chemical potential decreases with temperature (negative slope), and the slope magnitude increases from solid → liquid → gas because molar entropy increases.
  • Pressure effects: reducing pressure lowers chemical potential of all phases, but the gas phase shifts much more than solid or liquid due to its large molar volume.
  • Clapeyron equation: relates the slope of a coexistence curve (dp/dT) to measurable transition properties (entropy change or enthalpy change and volume change).
  • Common confusion: coexistence curve slopes—sublimation and vaporization curves always have positive slopes, but fusion (melting) curves can be positive or negative depending on whether volume increases or decreases on melting.

📉 Chemical potential and temperature

📉 How chemical potential changes with temperature

At constant pressure, the relationship is: (∂μ/∂T) at constant p = –S_m

  • The slope of chemical potential versus temperature at constant pressure is always negative because molar entropy S_m is always positive.
  • This means chemical potential decreases as temperature increases (at fixed pressure).

🔄 Different slopes for different phases

  • The magnitude of the slope increases going from solid → liquid → gas.
  • Why: molar entropies of sublimation and vaporization are positive, meaning gas has higher entropy than liquid, and liquid has higher entropy than solid.
  • Example: in the H₂O curves at 0.03 bar, the solid curve has the smallest (least negative) slope, liquid has a steeper slope, and gas has the steepest slope.

⚖️ Phase stability and intersections

  • At any given temperature and pressure, the stable phase is the one with the lowest chemical potential.
  • Where two curves intersect: the two phases coexist in equilibrium (e.g., point A is the melting point where solid and liquid curves cross; point B is the boiling point where liquid and gas curves cross).

🔽 Pressure effects on chemical potential

🔽 How pressure changes chemical potential

At constant temperature, the relationship is: (∂μ/∂p) at constant T = V_m

  • Reducing pressure at constant temperature lowers the chemical potential of every phase.
  • The downward shift is proportional to the molar volume of the phase.

🎈 Why gas shifts more than solid or liquid

  • Gas has a much larger molar volume than solid or liquid.
  • Example: when H₂O pressure drops from 0.03 bar to 0.003 bar:
    • Solid and liquid curves shift down by only ~0.002 kJ/mol (too small to see on the graph).
    • Gas curve shifts substantially downward, enough to intersect the solid curve at a sublimation point (point C).

🧊 Below the triple-point pressure

  • At pressures below the triple-point pressure (0.0062 bar for H₂O), only solid–gas equilibrium is possible.
  • The liquid phase is not stable at any temperature below the triple-point pressure.
  • Example: at 0.003 bar, the gas curve intersects only the solid curve (point C), not the liquid curve.

📐 The Clapeyron equation

📐 Deriving the coexistence condition

  • Start with two coexisting phases α and β at equilibrium: their chemical potentials are equal.
  • If temperature changes by dT, chemical potentials change unequally (because phases have different molar entropies).
  • To maintain equilibrium, pressure must also change by dp such that: dμ_α = dμ_β
  • Using dμ = –S_m dT + V_m dp for each phase and rearranging gives the Clapeyron equation.

📐 Two forms of the Clapeyron equation

Entropy form: dp/dT = Δ_trs S / Δ_trs V

Enthalpy form: dp/dT = Δ_trs H / (T Δ_trs V)

  • Both forms contain no approximations and apply to any phase transition of a pure substance.
  • The slope dp/dT of the coexistence curve depends on measurable quantities: transition enthalpy, entropy, and volume change.
  • The two forms are equivalent because Δ_trs S = Δ_trs H / T_trs (from earlier thermodynamic relations).

➕ Positive slopes: sublimation and vaporization

  • For sublimation (solid → gas) and vaporization (liquid → gas):
    • Both Δ_trs H and Δ_trs V are positive (enthalpy increases, volume increases).
    • Therefore, dp/dT is positive: the coexistence curve slopes upward.
  • Example: the solid–gas and liquid–gas coexistence curves in the H₂O phase diagram both slope upward to the right.

➕➖ Positive or negative slope: fusion (melting)

  • For fusion (solid → liquid):
    • Δ_fus H is always positive (melting requires heat input).
    • Δ_fus V may be positive or negative depending on the substance (most substances expand on melting, but some like H₂O contract).
  • Therefore, the solid–liquid coexistence curve slope depends on the sign of Δ_fus V.
  • Don't confuse: the sign of the fusion curve slope is substance-specific, unlike sublimation and vaporization curves which are always positive.

🗺️ Phase diagrams and coexistence curves

🗺️ Constructing the phase diagram

  • The coexistence curve is the projection of the intersection of two chemical potential surfaces onto the p–T plane.
  • Example: the heavy curve where liquid and gas chemical potential surfaces intersect (in 3D) projects down to the liquid–gas coexistence curve in the p–T phase diagram.

🔺 The triple point

  • The triple-point pressure is the unique pressure where all three phases (solid, liquid, gas) can coexist.
  • For H₂O, the triple-point pressure is 0.0062 bar.
  • Above this pressure: solid–liquid and liquid–gas coexistence are possible (points A and B at 0.03 bar).
  • Below this pressure: only solid–gas coexistence is possible (point C at 0.003 bar); liquid is never stable.

📊 Summary of H₂O phase behavior

PressureStable phasesCoexistence points
0.03 bar (above triple point)Solid, liquid, gasA: solid–liquid at 273.16 K; B: liquid–gas at 297.23 K
0.003 bar (below triple point)Solid, gas onlyC: solid–gas at 264.77 K; liquid never stable
  • The table shows how pressure determines which phase transitions are possible.
  • Don't confuse: at pressures below the triple point, there is no liquid phase at any temperature—the substance goes directly from solid to gas (sublimation).
48

Phase Equilibria

8.1 Phase Equilibria

🧭 Overview

🧠 One-sentence thesis

The Clapeyron equation quantifies how the slope of a phase coexistence curve depends on the enthalpy and volume changes of the transition, explaining why different phase boundaries have positive or negative slopes.

📌 Key points (3–5)

  • What the Clapeyron equation describes: the slope of a coexistence curve (d p / d T) in terms of measurable transition properties (enthalpy and volume changes).
  • Why phases remain in equilibrium: when temperature changes, pressure must also change so that the chemical potentials of both phases change equally.
  • Sublimation and vaporization curves: always have positive slopes because both enthalpy change and volume change are positive.
  • Fusion (melting) curves: can have positive or negative slopes depending on whether volume increases or decreases upon melting.
  • Common confusion: the sign of the slope depends on the sign of the volume change—most substances expand on melting (positive slope), but some contract (negative slope, like H₂O).

🔄 Equilibrium conditions during temperature and pressure changes

🔄 Maintaining phase equilibrium

  • When two phases (α' and α") coexist in equilibrium, their chemical potentials are equal.
  • If temperature changes by d T without pressure change, the chemical potentials change unequally → phases fall out of equilibrium.
  • To keep phases in equilibrium during a temperature change d T, there must be a simultaneous pressure change d p.

⚖️ Equal chemical potential changes

The requirement is:

d μ' = d μ"

  • The infinitesimal change of chemical potential in a phase is: d μ = −S_m d T + V_m d p
  • For two phases to remain in equilibrium:
    • −S'_m d T + V'_m d p = −S"_m d T + V"_m d p
  • Rearranging gives the relationship between d p and d T that maintains equilibrium.

📐 The Clapeyron equation

📐 Derivation and forms

Rearranging the equilibrium condition yields:

d p / d T = (S"_m − S'_m) / (V"_m − V'_m)

Which can be written as:

Clapeyron equation (form 1): d p / d T = Δ_trs S / Δ_trs V

Where:

  • Δ_trs S = molar entropy change of the transition
  • Δ_trs V = molar volume change of the transition

An alternative form uses the relationship Δ_trs S = Δ_trs H / T_trs:

Clapeyron equation (form 2): d p / d T = Δ_trs H / (T Δ_trs V)

🎯 What it tells us

  • The Clapeyron equation contains no approximations.
  • It gives the slope of the coexistence curve (the boundary between two phases on a pressure-temperature phase diagram).
  • The slope depends on quantities that can be measured: enthalpy change, volume change, and temperature.

🔺 Predicting coexistence curve slopes

🔺 Sublimation and vaporization curves

For solid → gas (sublimation) and liquid → gas (vaporization):

  • Both Δ_trs H and Δ_trs V are positive (the gas phase has higher enthalpy and much larger volume).
  • According to the Clapeyron equation, d p / d T is positive.
  • Implication: solid–gas and liquid–gas coexistence curves always slope upward (pressure increases with temperature).

🔻 Fusion (melting) curves

For solid → liquid (fusion):

  • Δ_fus H is always positive (melting requires energy input).
  • Δ_fus V may be positive or negative depending on the substance:
    • Most substances expand on melting → Δ_fus V > 0 → positive slope
    • Some substances (like H₂O) contract on melting → Δ_fus V < 0 → negative slope
  • Don't confuse: the sign of the slope is determined by the sign of the volume change, not the enthalpy change.
TransitionΔ_trs HΔ_trs VSlope sign
Sublimation+++ (always)
Vaporization+++ (always)
Fusion++ or −+ or − (substance-dependent)

🧊 Example: H₂O phase diagram

The excerpt references Figure 8.13(b) showing H₂O's pressure-temperature phase diagram:

  • At p = 0.03 bar: solid and liquid coexist at T = 273.16 K (point A); liquid and gas coexist at T = 297.23 K (point B).
  • At p = 0.003 bar (below triple-point pressure): only solid and gas coexist at T = 264.77 K (point C); the liquid phase is not stable.
  • The solid–liquid coexistence curve for H₂O has a negative slope because ice has a larger volume than liquid water (Δ_fus V < 0).
49

8.2 Phase Diagrams of Pure Substances

8.2 Phase Diagrams of Pure Substances

🧭 Overview

🧠 One-sentence thesis

The Clapeyron equation quantifies the slope of coexistence curves on phase diagrams by relating pressure–temperature changes to measurable transition properties, revealing why different phase boundaries have different slopes.

📌 Key points (3–5)

  • What the Clapeyron equation does: gives the slope (d p / d T) of coexistence curves using measurable quantities like transition enthalpy and volume change.
  • Sublimation and vaporization curves: always have positive slopes because both enthalpy change and volume change are positive.
  • Fusion (melting) curves: can have positive or negative slopes depending on whether the substance expands or contracts on melting.
  • Common confusion: most substances expand on melting (positive slope for solid–liquid curve), but water and a few others contract on melting (negative slope).
  • Why slopes matter: the steepness and direction of coexistence curves reflect the physical properties of phase transitions.

📐 The Clapeyron equation

📐 Two equivalent forms

The excerpt presents the Clapeyron equation in two forms:

Form 1 (Eq. 8.4.4): d p / d T = Δ_trs S / Δ_trs V
Form 2 (Eq. 8.4.5): d p / d T = Δ_trs H / (T Δ_trs V)

  • Both forms contain no approximations and apply to pure substances.
  • Form 2 substitutes the relationship Δ_trs S = Δ_trs H / T_trs into Form 1.
  • What it calculates: the slope of the coexistence curve (how pressure changes with temperature along a phase boundary).
  • Why it's useful: the slope can be determined from measurable quantities—transition enthalpy (Δ_trs H), transition volume change (Δ_trs V), and temperature.

🔬 Derivation context

  • The equation is derived by rearranging expressions involving chemical potentials of coexisting phases.
  • At coexistence, two phases have equal chemical potential; the Clapeyron equation describes how this equilibrium shifts with temperature and pressure.

🌡️ Sublimation and vaporization curves

🌡️ Always positive slopes

  • For sublimation (solid → gas) and vaporization (liquid → gas), both Δ_trs H and Δ_trs V are positive.
  • According to Eq. 8.4.5, when both numerator and denominator are positive, d p / d T is positive.
  • Implication: solid–gas and liquid–gas coexistence curves slope upward (pressure increases with temperature).

💨 Physical reasoning

  • Sublimation and vaporization both absorb heat (positive enthalpy change).
  • Both transitions involve expansion into the gas phase (positive volume change).
  • Example: On a phase diagram, the liquid–gas boundary rises as you move to higher temperatures, requiring higher pressures to keep the liquid phase stable.

🧊 Fusion (melting) curves

🧊 Variable slope direction

  • For fusion (solid → liquid), Δ_fus H is always positive (melting absorbs heat).
  • However, Δ_fus V may be positive or negative depending on the substance.
  • Result: the solid–liquid coexistence curve may have either a positive or negative slope.

📏 Steepness of the curve

  • The absolute value of Δ_fus V is small compared to vaporization or sublimation.
  • This makes the solid–liquid coexistence curve relatively steep (large pressure change for small temperature change).
  • The excerpt notes that in Fig. 8.13(b) and Fig. 8.2(c) for CO₂, the curve is so steep it's difficult to see the positive slope.

🔄 Expansion vs contraction on melting

🔄 Most substances expand

  • Most substances expand on melting (Δ_fus V > 0), giving a positive slope for the solid–liquid curve.
  • Example: Carbon dioxide expands on melting, so its solid–liquid boundary has a positive slope (though very steep).

❄️ Exceptions that contract

The excerpt lists substances that contract on melting at ordinary pressures:

SubstanceBehavior on melting
H₂O (water)Contracts (Δ_fus V < 0)
Rubidium nitrateContracts
AntimonyContracts
BismuthContracts
GalliumContracts
  • For these substances, Δ_fus V is negative, so the solid–liquid coexistence curve has a negative slope.
  • Don't confuse: water's unusual behavior (ice is less dense than liquid water) is reflected in its phase diagram as a negative-slope fusion curve, unlike most other substances.

🧊 Water's phase diagram

  • The excerpt mentions that the phase diagram for H₂O illustrates this exception.
  • Points A, B, and C in the figure correspond to different coexistence conditions at different pressures (0.03 bar and 0.003 bar).
  • At point A (273.16 K, 0.03 bar), solid and liquid coexist; at point B (297.23 K, 0.03 bar), liquid and gas coexist; at point C (264.77 K, 0.003 bar), solid and gas coexist.

📊 Interpreting phase diagrams

📊 Coexistence curves and slopes

  • Each coexistence curve represents conditions where two phases are in equilibrium.
  • The slope of each curve is determined by the signs and magnitudes of Δ_trs H and Δ_trs V for that transition.
  • Practical use: knowing the slope helps predict how phase boundaries shift with temperature and pressure changes.

🔍 Pressure effects

  • The excerpt notes that the effect of pressure on the chemical potential curves for solid and liquid is negligible compared to the gas phase.
  • Gas chemical potentials are shown at two different pressures (0.03 bar and 0.003 bar) to illustrate how pressure shifts equilibrium.
  • Example: At lower pressure (0.003 bar), the solid–gas coexistence occurs at a lower temperature (264.77 K) than the solid–liquid coexistence at higher pressure (273.16 K at 0.03 bar).
50

Phase Transitions

8.3 Phase Transitions

🧭 Overview

🧠 One-sentence thesis

The Clapeyron and Clausius–Clapeyron equations describe how coexistence curves between phases depend on temperature and pressure, enabling prediction of phase equilibria from measurable thermodynamic quantities.

📌 Key points (3–5)

  • Calorimetric measurement: Electrical work in an adiabatic calorimeter measures transition enthalpies by quantifying the substance transferred between coexisting phases.
  • Standard vs. real transitions: Standard molar quantities (with °) involve hypothetical ideal gas states and require multi-step calculations, while non-standard quantities refer to real coexisting phases in equilibrium.
  • Clapeyron equation: The exact slope of any coexistence curve is d p / d T = Δ_trs H / (T Δ_trs V), valid for all phase transitions without approximation.
  • Clausius–Clapeyron equation: For vaporization or sublimation (not near the critical point), approximating the gas as ideal gives d ln(p/p°) / d(1/T) = −Δ_trs H / R.
  • Common confusion: The solid–liquid curve is usually steep and positive (most substances expand on melting), but water contracts on melting, giving a negative slope for ice I–liquid.

🔬 Measuring transition enthalpies

⚡ Electrical work method

The most precise measurement of the molar enthalpy of an equilibrium phase transition uses electrical work.

  • A constant-pressure adiabatic calorimeter contains coexisting phases.
  • A known quantity of electrical work (I² R_el Δt) is performed on the system.
  • The amount of substance transferred between phases is measured.
  • The first law shows that this electrical work equals the heat that would be needed for the same change of state.
  • At constant pressure, this heat equals the enthalpy change of the process.
  • Unlike heat capacity measurements, temperature remains constant and no calorimeter heat capacity correction is needed.

📐 Standard molar transition quantities

🎯 What "standard" means

Standard molar enthalpy of vaporization, Δ_vap H°, is the enthalpy change when pure liquid in its standard state at a specified temperature changes to gas in its standard state at the same temperature, divided by the amount changed.

  • The initial state (pure liquid at pressure p°) is real.
  • The final state (gas behaving ideally at pressure p°) is hypothetical.
  • The liquid and gas are not necessarily in equilibrium with one another at p° and the temperature of interest.

🔄 Standard vs. non-standard transitions

QuantitySymbolStates involvedDirect calorimetric measurement?
Standard vaporizationΔ_vap H°Real liquid → hypothetical ideal gasNo, requires corrections
Standard sublimationΔ_sub H°Real solid → hypothetical ideal gasNo, requires corrections
Non-standard vaporizationΔ_vap HReal liquid → real gas (coexisting)Yes, reversible transition
Non-standard sublimationΔ_sub HReal solid → real gas (coexisting)Yes, reversible transition
Standard fusionΔ_fus H°Real solid → real liquid (coexisting)Yes, reversible transition

Don't confuse: Δ_vap H° and Δ_sub H° cannot be measured directly from coexisting phases because the phases are not in equilibrium at standard pressure.

🛤️ Three-step calculation path

Standard molar transition quantities (Δ_vap X° or Δ_sub X°) are functions only of temperature T. To evaluate them:

  1. Isothermal pressure change of condensed phase: From standard state at p° to vapor pressure p_vap at temperature T.
    • Use expressions from Table 7.4 (mentioned in excerpt).
  2. Reversible vaporization or sublimation: At T and p_vap, forming real gas.
    • This step gives Δ_vap X or Δ_sub X, which can be measured experimentally.
  3. Isothermal change of real gas: From p_vap to hypothetical ideal gas at p°.
    • Use formulas from Table 7.5 relating real gas to standard quantities.

The sum of ΔX_m for these three steps equals the desired Δ_vap X° or Δ_sub X°.

🗺️ Chemical potential surfaces and coexistence

📊 How phases coexist

  • Chemical potential μ of a pure substance in a single phase is a function of T and p, represented as a three-dimensional surface.
  • Equilibrium condition: Two phases coexist only when they have the same T, p, and μ.
  • Equilibrium exists only along the intersection of the two phase surfaces.
  • The projection of this intersection onto the p–T plane generates the coexistence curve.

📉 How μ varies with temperature

The partial derivatives of chemical potential are:

  • (∂μ/∂T)_p = −S_m
  • (∂μ/∂p)_T = V_m

At constant pressure, μ versus T behavior:

  • The slope is always negative (since molar entropy is always positive).
  • The magnitude of the slope increases going from solid → liquid → gas (because molar entropies of sublimation and vaporization are positive).
  • The stable phase at each temperature is the one with the lowest μ (transfer from higher to lower μ at constant T and p is spontaneous).

Example: For H₂O at 0.03 bar (above triple-point pressure):

  • Solid and liquid curves intersect at a melting point (point A).
  • Liquid and gas curves intersect at a boiling point (point B).

🔽 Effect of pressure reduction

  • From (∂μ/∂p)_T = V_m, a pressure reduction at constant temperature lowers the chemical potential of a phase.
  • The shift is proportional to the molar volume of the phase.
  • For condensed phases (solid, liquid), the shift is very small (Δμ ≈ −0.002 kJ/mol for H₂O from 0.03 to 0.003 bar).
  • For gas, the shift is substantial because of large molar volume.

Example: For H₂O at 0.003 bar (below triple-point pressure):

  • The gas curve shifts downward significantly and now intersects the solid curve at a sublimation point (point C).
  • Liquid phase is not stable at any pressure below the triple-point pressure.
  • Only solid–gas equilibrium is possible.

🧮 The Clapeyron equation

🔧 Derivation from equilibrium condition

Starting with two coexisting phases α and β:

  • If temperature changes by dT without pressure change, phases are no longer in equilibrium (chemical potentials change unequally).
  • For phases to remain in equilibrium during temperature change dT, there must be a simultaneous pressure change dp such that dμ^α = dμ^β.
  • Since dμ = −S_m dT + V_m dp, the condition becomes:
    • −S^α_m dT + V^α_m dp = −S^β_m dT + V^β_m dp

Rearranging gives:

Clapeyron equation: d p / d T = Δ_trs S / Δ_trs V (pure substance)

Or equivalently:

d p / d T = Δ_trs H / (T Δ_trs V) (pure substance)

(using Δ_trs S = Δ_trs H / T_trs)

Key feature: Contains no approximations; valid for all phase transitions.

📈 Predicting coexistence curve slopes

TransitionΔ_trs HΔ_trs VSlope sign
Sublimation (solid → gas)PositivePositivePositive
Vaporization (liquid → gas)PositivePositivePositive
Fusion (solid → liquid)PositivePositive or negativeEither

For fusion:

  • Δ_fus H is always positive.
  • Δ_fus V may be positive (expansion on melting) or negative (contraction on melting).
  • The absolute value of Δ_fus V is small, making the solid–liquid coexistence curve relatively steep.

❄️ Special cases for fusion

Most substances expand on melting (positive Δ_fus V):

  • Solid–liquid coexistence curve has positive slope.
  • Example: Carbon dioxide (though the curve is so steep the positive slope is hard to see).

Exceptions that contract on melting (negative Δ_fus V):

  • H₂O, rubidium nitrate, antimony, bismuth, gallium.
  • Solid–liquid coexistence curve has negative slope.
  • Example: Ice I and liquid water—ordinary ice is less dense than liquid water, giving a negative slope in the phase diagram.
  • High-pressure forms of ice are more dense than liquid, so those solid–liquid curves have positive slopes.
  • Ice VII–Ice VIII curve is vertical (identical molar volumes because they have identical crystal structures except for H₂O molecule orientations).

🧊 Melting point dependence on pressure

For small changes of T and p, if Δ_fus V and Δ_fus H are approximately constant, integrating the Clapeyron equation gives:

  • p₂ − p₁ ≈ (Δ_fus H / Δ_fus V) ln(T₂ / T₁)

Or:

  • T₂ ≈ T₁ exp[Δ_fus V (p₂ − p₁) / Δ_fus H]

This estimates how melting point depends on pressure.

🌡️ The Clausius–Clapeyron equation

🎈 Approximations for gas–condensed phase equilibria

When gas coexists with liquid or solid, provided T and p are not close to the critical point:

  1. Molar volume approximation: The gas molar volume is much greater than the condensed phase:

    • Δ_vap V = V^g_m − V^l_m ≈ V^g_m
    • Δ_sub V = V^g_m − V^s_m ≈ V^g_m
  2. Ideal gas approximation: The gas behaves as an ideal gas:

    • V^g_m ≈ RT / p

Substituting into the Clapeyron equation gives:

Clausius–Clapeyron equation: d p / d T ≈ p Δ_trs H / (R T²) (pure substance, vaporization or sublimation)

Don't confuse: This is not valid for:

  • Coexisting solid and liquid phases.
  • Coexisting liquid and gas phases close to the critical point.

🔗 Triple point relation

At the triple point, all three transitions are possible:

  • Fusion followed by vaporization equals sublimation.
  • Therefore: Δ_fus H + Δ_vap H = Δ_sub H

Since all three are positive:

  • Δ_sub H > Δ_vap H at the triple point.
  • According to the Clausius–Clapeyron equation, the slope of the solid–gas curve at the triple point is slightly greater than the slope of the liquid–gas curve.

📊 Alternative forms for vapor pressure

Rearranging the Clausius–Clapeyron equation:

  1. Form 1: d ln(p/p°) / d T ≈ Δ_trs H / (R T²)

  2. Form 2: d ln(p/p°) ≈ −(Δ_trs H / R) d(1/T)

  3. Form 3: d ln(p/p°) / d(1/T) ≈ −Δ_trs H / R

Practical use: A plot of ln(p/p°) versus 1/T for vapor pressure data gives a curve whose slope at each temperature equals −Δ_vap H / R or −Δ_sub H / R at that temperature (usually to high accuracy).

Example: This allows determination of transition enthalpies from vapor pressure measurements over a range of temperatures.

51

Coexistence Curves

8.4 Coexistence Curves

🧭 Overview

🧠 One-sentence thesis

The Clapeyron and Clausius–Clapeyron equations quantitatively describe how pressure and temperature vary along phase coexistence curves, enabling prediction of melting points, boiling points, and transition enthalpies from experimental data.

📌 Key points (3–5)

  • Clapeyron equation: relates the slope of any coexistence curve (dp/dT) to the transition enthalpy and volume change.
  • Clausius–Clapeyron equation: a simplified form for liquid–gas or solid–gas equilibria that assumes the gas volume dominates and behaves ideally.
  • When to use which: Clausius–Clapeyron is valid only for vaporization/sublimation away from the critical point; Clapeyron applies to all phase transitions including solid–liquid.
  • Common confusion: the Clausius–Clapeyron equation does NOT work for solid–liquid coexistence or near the critical point, where the gas volume approximation breaks down.
  • Practical application: plotting ln(p/p°) versus 1/T yields a straight line whose slope gives the transition enthalpy without calorimetry.

📐 The Clapeyron equation

📐 General form for any phase transition

The Clapeyron equation: dp/dT = Δ_trs H / (T Δ_trs V), where Δ_trs H is the molar enthalpy of transition and Δ_trs V is the molar volume change.

  • This equation applies to any phase coexistence curve (solid–liquid, liquid–gas, solid–gas).
  • It describes how pressure must change with temperature to maintain equilibrium between two phases.
  • The slope depends on both the enthalpy change and the volume change of the transition.

🧊 Example: ice and water coexistence

  • The excerpt notes that ice I has a negative slope on its coexistence curve with liquid water.
  • Reason: ordinary ice is less dense than liquid water, so Δ_fus V is negative.
  • High-pressure forms of ice are denser than liquid, so their solid–liquid curves have positive slopes.
  • Ice VII and ice VIII have a vertical coexistence curve because they have identical molar volumes (same crystal structure, different molecular orientations).

🔄 Rearranged form for solid–liquid transitions

The Clapeyron equation can be rearranged and integrated:

  • If Δ_fus V and Δ_fus H are approximately constant over small changes in T and p, integration gives:
    • p₂ − p₁ = (Δ_fus H / Δ_fus V) ln(T₂/T₁)
    • Or: T₂ = T₁ exp[(Δ_fus V (p₂ − p₁)) / Δ_fus H]
  • This allows estimation of how the melting point depends on pressure.
  • Why the approximation works: cubic expansion coefficient and isothermal compressibility of condensed phases are relatively small.

🌡️ The Clausius–Clapeyron equation

🌡️ Simplification for gas–condensed phase equilibria

The Clausius–Clapeyron equation: dp/dT = p Δ_trs H / (R T²), valid for vaporization or sublimation when the gas volume dominates and behaves ideally.

  • Key assumptions:
    1. The molar volume of the gas is much greater than that of the condensed phase: Δ_vap V ≈ V_g^m and Δ_sub V ≈ V_g^m.
    2. The gas behaves as an ideal gas: V_g^m = RT/p.
  • Not valid for:
    • Solid–liquid coexistence.
    • Liquid–gas coexistence close to the critical point (where gas and liquid volumes become comparable).

🔬 Alternative forms for practical use

The Clausius–Clapeyron equation can be rewritten in several equivalent forms:

FormExpressionUse
Basicdp/dT = p Δ_trs H / (R T²)Shows slope of coexistence curve
Logarithmicd ln(p/p°) / dT = Δ_trs H / (R T²)Easier to integrate
Inverse temperatured ln(p/p°) / d(1/T) = −Δ_trs H / RLinear plot form
  • The last form shows that a plot of ln(p/p°) versus 1/T has a slope equal to −Δ_vap H / R or −Δ_sub H / R.
  • This provides an alternative to calorimetry for measuring transition enthalpies.

📊 Pressure units and standard pressure

  • The ratio p/p° is dimensionless, where p° is standard pressure (1 bar).
  • For evaluating Δ_trs H, any convenient pressure unit can be used: ln(p/bar), ln(p/Pa), or ln(p/Torr) all give the same slope (but different intercepts).
  • The numerical value of p/p° is simply the pressure value when expressed in the chosen units.

🔺 At the triple point

🔺 Relationship between transition enthalpies

At the triple point, all three phases (solid, liquid, gas) coexist, and the transition enthalpies are related:

Δ_fus H + Δ_vap H = Δ_sub H

  • This means: fusion followed by vaporization equals sublimation.
  • Since all three transition enthalpies are positive, it follows that Δ_sub H > Δ_vap H at the triple point.

📈 Slopes of coexistence curves

  • According to the Clausius–Clapeyron equation, the slope of a coexistence curve is proportional to Δ_trs H.
  • Therefore, at the triple point, the solid–gas curve has a slightly greater slope than the liquid–gas curve.
  • This is because Δ_sub H > Δ_vap H.

🧮 Integrated form and estimation

🧮 Constant enthalpy approximation

If Δ_vap H or Δ_sub H is essentially constant over a temperature range, the Clausius–Clapeyron equation can be integrated:

ln(p₂/p₁) = (Δ_trs H / R) (1/T₁ − 1/T₂)

  • This equation allows estimation of any one of the five quantities (p₁, p₂, T₁, T₂, or Δ_trs H) given the other four.
  • Example: knowing the vapor pressure at one temperature and the enthalpy of vaporization, you can estimate the vapor pressure at another temperature.

⚠️ Don't confuse: when integration is valid

  • The integrated form assumes Δ_trs H is constant over the temperature range.
  • In reality, transition enthalpies vary with temperature, so this is an approximation.
  • The approximation is better for smaller temperature ranges.
  • For more accurate results over larger ranges, use the differential form with temperature-dependent Δ_trs H.
52

Composition Variables

9.1 Composition Variables

🧭 Overview

🧠 One-sentence thesis

Composition variables are intensive properties that quantify the relative amount of each species in a mixture, and they differ in how they respond to changes in volume or pressure.

📌 Key points (3–5)

  • What composition variables are: intensive properties indicating the relative amount of a particular species in a phase.
  • Four main types: mole fraction, mass fraction, concentration (molarity), and molality (for solutes in solutions).
  • Species vs substance distinction: a species is any entity with definite composition and charge (e.g., ions), but only uncharged or neutral species that can be prepared pure are substances.
  • Common confusion—fixed composition: when composition is "fixed" or "constant," mole fractions, mass fractions, and molalities remain constant, but concentrations and partial pressures may change if volume or pressure changes.
  • Dilute solution limit: in infinitely dilute solutions, all solute composition variables (mole fraction, mass fraction, concentration, molality) become proportional to one another.

🧩 Core concepts

🧩 What is a composition variable

A composition variable is an intensive property that indicates the relative amount of a particular species or substance in a phase.

  • It is not an absolute amount (like moles or mass), but a relative measure.
  • Because it is intensive, it does not depend on the total size of the system.
  • Example: two identical mixtures of different total amounts have the same composition variable values.

🔬 Species vs substance

A species is any entity of definite elemental composition and charge, described by a chemical formula (e.g., H₂O, H₃O⁺, NaCl, Na⁺).

A substance is a species that can be prepared in a pure state (e.g., N₂ and NaCl).

  • Charged species (ions) like H₃O⁺ or Na⁺ are species but not substances, because a macroscopic amount of a single kind of ion cannot be prepared by itself.
  • Don't confuse: every substance is a species, but not every species is a substance.

📊 Types of composition variables

📊 Mole fraction

  • Defined by: mole fraction of species i = (amount of species i) / (sum of amounts of all species).
  • Symbol: x_i for mixtures in general; y_i when the mixture is a gas.
  • The sum of all mole fractions equals 1.
  • Example: in a mixture with 2 mol of species A and 3 mol of species B, x_A = 2/(2+3) = 0.4.

⚖️ Mass fraction (weight fraction)

  • Defined by: mass fraction of species i = (mass of species i) / (total mass of mixture).
  • Symbol: w_i.
  • Expressed as: w_i = (n_i × M_i) / (sum over all j of n_j × M_j), where M is molar mass.
  • The sum of all mass fractions equals 1.

🧪 Concentration (molarity)

  • Defined by: concentration of species i = (amount of species i) / (volume of mixture).
  • Symbol: c_i; alternative notation: [A].
  • Units: mol L⁻¹ or mol dm⁻³, often written as "M" (molar).
  • Example: 0.5 M means 0.5 moles per liter.
  • Important: concentration changes if the volume changes, even if the relative amounts of species remain constant.

🧂 Molality (for solutes)

  • Defined by: molality of solute species B = (amount of solute B) / (mass of solvent A).
  • Symbol: m_B.
  • Units: mol kg⁻¹, sometimes written as "m" (molal), though this notation is discouraged because m also means meter.
  • Example: 0.6 molal means 0.6 moles of solute per kilogram of solvent.
  • Only used for solutes in solutions, not for the solvent or for general mixtures.

🧪 Solutions and their special features

🧪 What is a solution

A solution is a mixture in which one substance (the solvent, denoted A) is treated specially; all other species are solutes (denoted B, C, etc.).

  • Although a solution can in principle be a gas mixture, this section considers only liquid and solid solutions.
  • A solution can be prepared by gradually mixing solutes with the solvent while maintaining the same physical state (liquid or solid) as the pure solvent.
  • Dilute solution: the sum of solute mole fractions is small compared to the solvent mole fraction (x_A is close to 1).
  • Concentrated solution: as solute mole fractions increase, the solution becomes more concentrated.

🔗 Binary solutions

A binary mixture (or binary solution) is a mixture of exactly two substances.

  • For a binary solution of solvent A and solute B, simplified equations relate the composition variables.
  • The mole ratio n_B / n_A can be expressed in terms of any composition variable:
    • From mole fraction: n_B / n_A = x_B / (1 − x_B)
    • From mass fraction: n_B / n_A = (M_A × w_B) / [M_B × (1 − w_B)]
    • From concentration: n_B / n_A = (M_A × c_B) / (ρ − M_B × c_B), where ρ is solution density
    • From molality: n_B / n_A = M_A × m_B
  • These expressions allow conversion between different composition variables.

♾️ Infinite dilution limit

  • In the limit of infinite dilution (x_A → 1), all solute composition variables become proportional to one another:
    • n_B / n_A = x_B = (M_A / M_B) × w_B = (M_A / ρ*_A) × c_B = M_A × m_B
    • ρ*_A is the density of pure solvent A.
  • This proportionality also holds for any solute B in a multi-solute solution where each solute is very dilute.
  • Example: for dilute aqueous solutions, molarity (c_B in mol L⁻¹) ≈ molality (m_B in mol kg⁻¹), because the density of water is approximately 1 kg L⁻¹.

🔄 Fixed vs changing composition

🔄 What "fixed composition" means

  • When composition is described as fixed or constant during changes of temperature, pressure, or volume, it means there is no change in the relative amounts or masses of the species.
  • A mixture of fixed composition has fixed values of:
    • Mole fractions
    • Mass fractions
    • Molalities
  • But not necessarily fixed values of:
    • Concentrations (will change if volume changes)
    • Partial pressures in a gas mixture (will change if pressure changes)

⚠️ Don't confuse

  • "Fixed composition" does not mean all composition variables stay constant.
  • It means the relative amounts stay constant, so only amount-based or mass-based ratios remain unchanged.
  • Example: if a gas mixture is compressed at constant temperature, the partial pressures increase even though the composition (mole fractions) is fixed.

📐 Partial molar quantities (introduction)

📐 Definition

The partial molar quantity X_i of species i is defined by: X_i = (∂X / ∂n_i) at constant T, p, and amounts of all other species (n_j≠i).

  • X is an extensive property of the mixture (e.g., volume, enthalpy).
  • X_i is the rate at which property X changes when species i is added to the mixture, holding temperature, pressure, and all other species' amounts constant.
  • A partial molar quantity is an intensive state function; its value depends on temperature, pressure, and composition.

⚡ Practical limitation for ions

  • If species i is charged (an ion), X_i as defined above is a theoretical concept.
  • A macroscopic amount of a charged species cannot be added by itself to a phase, because it would create a huge electric charge.
  • Therefore, partial molar quantities of ions cannot be determined experimentally in isolation.

🧊 Partial molar volume example

  • The partial molar volume V_B of species B is: V_B = (∂V / ∂n_B) at constant T, p, and n_A.
  • It represents the rate at which the system volume changes when methanol (substance B) is added to a water–methanol mixture at constant temperature and pressure.
  • Example (from figure description): adding one mole (40.75 cm³) of pure methanol to a water–methanol mixture increases the mixture volume by only 38.8 cm³, showing that the partial molar volume of methanol in the mixture is less than the molar volume of pure methanol.
53

Partial Molar Quantities

9.2 Partial Molar Quantities

🧭 Overview

🧠 One-sentence thesis

Partial molar quantities measure how an extensive property of a mixture changes when a small amount of one component is added at constant temperature and pressure, providing a powerful framework for understanding mixture behavior and equilibrium.

📌 Key points (3–5)

  • Definition: A partial molar quantity X_i is the rate of change of extensive property X with the amount of species i added, holding T, p, and all other amounts constant.
  • Key insight: When you add one mole of a substance to a large mixture, the volume (or other property) change is usually not equal to the molar volume of the pure substance—intermolecular interactions in the mixture cause deviations.
  • Additivity rule: The total value of any extensive property X equals the sum of (amount of each species) × (its partial molar quantity).
  • Gibbs–Duhem constraint: Changes in partial molar quantities are linked—if one increases with composition change, another must decrease in a related way.
  • Chemical potential: The partial molar Gibbs energy is called the chemical potential and governs the escaping tendency of a species, central to equilibrium problems.

📐 Defining partial molar quantities

📐 The core definition

Partial molar quantity X_i: the rate at which property X changes with the amount of species i added to the mixture, at constant T, p, and amounts of all other species: X_i = (∂X/∂n_i) at constant T, p, n_j≠i

  • X is any extensive property (volume, enthalpy, entropy, Gibbs energy, etc.).
  • The subscript i identifies a constituent species.
  • This is an intensive state function—its value depends on T, p, and composition, not on system size.

⚠️ Practical limitation for charged species

  • A partial molar quantity of an ion (charged species) is a theoretical concept.
  • You cannot physically add a macroscopic amount of a single ion to a phase without creating a huge electric charge.
  • Therefore, partial molar quantities of ions cannot be determined experimentally in isolation (though relative values can be found using reference ions).

🧪 Partial molar volume: a concrete example

🧪 Water–methanol mixing behavior

The excerpt uses volume because it is easily visualized:

  • Pure water at 25 °C, 1 bar: molar volume = 18.07 cm³/mol
  • Pure methanol at 25 °C, 1 bar: molar volume = 40.75 cm³/mol
  • Mix 100.0 cm³ water + 100.0 cm³ methanol → total volume = 193.1 cm³, not 200.0 cm³
  • The 6.9 cm³ "missing" volume arises from new intermolecular interactions in the mixture.

🔬 Adding methanol to a large mixture

Imagine a large volume (10,000 cm³) of water–methanol mixture with mole fraction x_B = 0.307 (B = methanol):

  • Add 40.75 cm³ (one mole) of pure methanol.
  • The mixture volume increases by only 38.8 cm³, not 40.75 cm³.
  • Because the added amount is small compared to the total, the composition barely changes (x_B increases by only 0.5%).
  • This volume change per mole added approximates the partial molar volume: V_B ≈ 38.8 cm³/mol at this composition.

Interpretation: The partial molar volume V_B is the volume increase per mole of B when B is mixed with such a large volume of mixture that the composition is not appreciably affected—or equivalently, the volume change per amount when an infinitesimal amount is mixed with a finite volume.

🔄 Limiting values

  • As x_B → 1 (pure B), the partial molar volume V_B approaches the molar volume of pure B: V_B(x_B=1) = V°_B.
  • As x_B → 0 (infinite dilution), V_B approaches a limiting value V∞_B, the partial molar volume at infinite dilution (each solute molecule surrounded only by solvent).

➖ Negative partial molar volumes

  • It is possible for a partial molar volume to be negative.
  • Example: Magnesium sulfate in dilute aqueous solution (molality < 0.07 mol/kg) has negative V_MgSO₄.
  • Physical meaning: when a small amount of crystalline MgSO₄ dissolves in water at constant T, the liquid phase contracts.
  • Cause: strong attractive water–ion interactions.

🧮 Mathematical framework

🧮 Total differential for an open binary mixture

For a binary mixture (substances A and B), the volume has four independent variables: T, p, n_A, n_B.

The total differential of V is: dV = (∂V/∂T) dT + (∂V/∂p) dp + (∂V/∂n_A) dn_A + (∂V/∂n_B) dn_B

Recognizing the partial derivatives:

  • (∂V/∂T) at constant p, n_A, n_B = αV (thermal expansion)
  • (∂V/∂p) at constant T, n_A, n_B = −κ_T V (compressibility)
  • (∂V/∂n_A) at constant T, p, n_B = V_A (partial molar volume of A)
  • (∂V/∂n_B) at constant T, p, n_A = V_B (partial molar volume of B)

So: dV = αV dT − κ_T V dp + V_A dn_A + V_B dn_B

At constant T and p: dV = V_A dn_A + V_B dn_B

➕ Additivity rule

Imagine continuously pouring pure A and pure B at constant rates into a stirred container at constant T and p, forming a mixture of increasing volume and constant composition.

Because T, p, and composition are constant, V_A and V_B remain constant during the process.

Integrating dV = V_A dn_A + V_B dn_B gives:

Additivity rule: V = n_A V_A + n_B V_B (binary mixture)

This allows calculating the mixture volume from the amounts and the appropriate partial molar volumes for the particular T, p, and composition.

Example: For water–methanol mixture with x_B = 0.307, given V_A = 17.74 cm³/mol and V_B = 38.76 cm³/mol:

  • n_A = 5.53 mol, n_B = 2.45 mol
  • V = (17.74)(5.53) + (38.76)(2.45) = 193.1 cm³ ✓

🔗 Gibbs–Duhem equation

Differentiating the additivity rule V = n_A V_A + n_B V_B: dV = V_A dn_A + V_B dn_B + n_A dV_A + n_B dV_B

But we already know dV = V_A dn_A + V_B dn_B at constant T and p.

These are consistent only if:

Gibbs–Duhem equation: n_A dV_A + n_B dV_B = 0 (constant T and p)

Equivalently: x_A dV_A + x_B dV_B = 0

Meaning: Changes in V_A and V_B are linked when composition changes at constant T and p.

Rearranging: dV_A = −(n_B/n_A) dV_B

  • A composition change that increases V_B (positive dV_B) must make V_A decrease.
  • The water–methanol data show this mirroring: a minimum in V_B at x_B ≈ 0.09 corresponds to a maximum in V_A (though attenuated because n_B/n_A is small there).

📊 Experimental determination: method of intercepts

📊 Mean molar volume plot

Plot V/n (the mean molar volume) versus mole fraction x_B.

At the composition of interest x°_B:

  1. Draw the tangent line to the curve at that point.
  2. The tangent intercepts the vertical axis at x_B = 0 at V_A.
  3. The tangent intercepts the vertical axis at x_B = 1 at V_B.

Derivation sketch:

  • From the additivity rule: V/n = V_A x_A + V_B x_B = V_A(1 − x_B) + V_B x_B = (V_B − V_A)x_B + V_A
  • Differentiating: d(V/n)/dx_B = V_B − V_A + [terms involving dV_A/dx_B and dV_B/dx_B]
  • Using the Gibbs–Duhem equation, the derivative simplifies to: d(V/n)/dx_B = V_B − V_A
  • The tangent line at x°_B has slope (V°_B − V°_A) and passes through the point with ordinate (V°_B − V°_A)x°_B + V°_A.
  • This line has intercepts V°_A at x_B = 0 and V°_B at x_B = 1.

📈 Molar volume of mixing plot (variant)

An alternative is to plot the molar volume of mixing: ΔV_m(mix) = ΔV(mix)/n = V/n − x_A V°_A − x_B V°_B

where V°_A and V°_B are the molar volumes of pure A and pure B.

On this plot, the tangent at composition x°_B has intercepts:

  • At x_B = 0: V_A − V°_A
  • At x_B = 1: V_B − V°_B

This variant directly shows deviations from ideal mixing (where partial molar volumes would equal pure-component molar volumes).

🔍 Don't confuse: partial molar volume vs. molar volume

  • Molar volume of pure B (V°_B): volume per mole of pure substance B.
  • Partial molar volume of B in a mixture (V_B): volume change per mole of B added to a large amount of mixture at a given composition.
  • These are equal only when x_B = 1 (pure B).
  • At other compositions, V_B ≠ V°_B due to intermolecular interactions in the mixture.

🌐 General relations for any property and any number of components

🌐 Generalized definitions

For any extensive property X and species i in a mixture:

Partial molar quantity: X_i = (∂X/∂n_i) at constant T, p, n_j≠i

Additivity rule: X = Σ_i n_i X_i

Gibbs–Duhem equation: Σ_i n_i dX_i = 0 (constant T and p)

Equivalently: Σ_i x_i dX_i = 0 (constant T and p)

These apply to mixtures where each species i can be:

  • A nonelectrolyte substance
  • An electrolyte substance (dissociated into ions)
  • An individual ionic species

Important: In the Gibbs–Duhem equation, mole fraction x_i must be based on the species considered present. For aqueous NaCl, you could treat it as {H₂O, NaCl} or as {H₂O, Na⁺, Cl⁻}, but the mole fractions must be consistent with that choice.

🧪 What can and cannot be measured

Can measure experimentally (for uncharged species):

  • V_i (partial molar volume)
  • C_p,i (partial molar heat capacity)
  • S_i (partial molar entropy)

Cannot measure absolute values:

  • U_i, H_i, A_i, G_i (partial molar internal energy, enthalpy, Helmholtz energy, Gibbs energy)
  • Reason: these involve the internal energy brought into the system by the species, and absolute internal energy cannot be evaluated.
  • However, we can measure differences like H_i − H°_i from calorimetric measurements of mixing enthalpies.
  • These quantities are still useful in theoretical relations.

⚡ Partial molar quantities of ions (relative values)

For a fully dissociated electrolyte M^(ν₊)X^(ν₋) in aqueous solution: V_B = ν₊ V₊ + ν₋ V₋

where V₊ and V₋ are the (unmeasurable) partial molar volumes of the cation and anion.

Convention for aqueous solutions:

  • Reference ion: H⁺
  • Set V∞_H⁺ = 0 (partial molar volume of H⁺ at infinite dilution)

Example chain:

  • Measure V∞_HCl = 17.82 cm³/mol
  • Then "conventional" V∞_Cl⁻ = V∞_HCl − V∞_H⁺ = 17.82 cm³/mol
  • Measure V∞_NaCl = 16.61 cm³/mol
  • Then conventional V∞_Na⁺ = V∞_NaCl − V∞_Cl⁻ = 16.61 − 17.82 = −1.21 cm³/mol

These are called "conventional" values because they depend on the arbitrary choice V∞_H⁺ = 0.

🔬 Partial specific quantities

🔬 Definition and use

Partial specific quantity: the partial molar quantity divided by the molar mass.

Example: partial specific volume v_B = V_B / M_B = (∂V/∂m_B) at constant T, p, m_A

where m_A and m_B are masses (not amounts).

Advantage: Can be evaluated without knowing the molar mass.

  • Used in sedimentation equilibrium experiments to determine molar mass.

Analogous relations:

  • Replace amounts by masses
  • Replace mole fractions by mass fractions
  • Replace partial molar quantities by partial specific quantities

Example additivity: V = Σ_i m_i v_i

Example Gibbs–Duhem: Σ_i w_i dv_i = 0 (where w_i is mass fraction)

Method of intercepts for binary mixture:

  • Plot specific volume v versus mass fraction w_B
  • Tangent at composition w°_B has intercepts v_A at w_B = 0 and v_B at w_B = 1
  • Variant: plot (v − w_A v°_A − w_B v°_B) versus w_B; intercepts are (v_A − v°_A) and (v_B − v°_B)

⚗️ Chemical potential and equilibrium

⚗️ Definition of chemical potential

Chemical potential μ_i: the partial molar Gibbs energy of species i in a mixture: μ_i = (∂G/∂n_i) at constant T, p, n_j≠i

  • If nonexpansion work coordinates exist, hold them constant too.
  • The chemical potential is a measure of the escaping tendency of the species from the phase.
  • Crucial role in equilibrium problems.
  • We cannot determine the absolute value of μ_i, but we can usually evaluate the difference between the value in a given state and a defined reference state.

📝 Gibbs fundamental equation for mixtures

For an open single-phase system with s different nonreacting species, there are 2 + s independent variables (T, p, and n_i for each species).

Gibbs fundamental equation: dG = −S dT + V dp + Σ_i μ_i dn_i

This generalizes the closed-system equation by adding a term for each species to allow composition to vary.

⚡ Chemical potentials of ions (theoretical concept)

Consider aqueous KCl solution. Constituents could be:

  • {H₂O, KCl}: amounts can be independently varied in practice
  • {H₂O, K⁺, Cl⁻}: amounts cannot be independently varied (must maintain electrical neutrality)

The chemical potential μ_K⁺ is defined as the rate at which G changes with amount of K⁺ added at constant T, p, and amount of Cl⁻.

  • This is a hypothetical process in which the net charge increases.
  • Therefore, μ_K⁺ is a valid but purely theoretical concept.

Both formulations are valid:

  • dG = −S dT + V dp + μ_H₂O dn_H₂O + μ_KCl dn_KCl
  • dG = −S dT + V dp + μ_H₂O dn_H₂O + μ_K⁺ dn_K⁺ + μ_Cl⁻ dn_Cl⁻

🔄 Extension to multiphase, multicomponent systems

The excerpt mentions (but does not detail) that the equilibrium derivation from Section 8.1.2 (single substance, multiple phases) extends to systems with:

  • Two or more homogeneous phases
  • Mixtures of nonreacting species
  • No internal partitions preventing transfer
  • Negligible gravity and external force fields

The system consists of a reference phase α⁰ and other phases labeled α ≠ α⁰, with species transferable between phases.

(The excerpt ends before completing this derivation.)

54

Gas Mixtures

9.3 Gas Mixtures

🧭 Overview

🧠 One-sentence thesis

In gas mixtures, each constituent's chemical potential depends on its partial pressure, with ideal gas mixtures obeying a simple logarithmic relation while real gas mixtures require fugacity corrections to account for intermolecular interactions.

📌 Key points (3–5)

  • Partial pressure definition: the product of a constituent's mole fraction and the total pressure of the gas phase; the sum of all partial pressures equals the total pressure (Dalton's Law).
  • Ideal gas mixture criterion: each constituent i obeys μᵢ = μᵢ°(g) + RT ln(pᵢ/p°), where partial molar properties (H, U, Cp) depend only on temperature, not composition or pressure.
  • Real gas corrections: fugacity fᵢ replaces partial pressure pᵢ in the chemical potential expression, accounting for deviations from ideal behavior due to intermolecular interactions.
  • Common confusion: partial pressure in an ideal gas mixture equals the pressure the pure substance would have at the same T and V only for ideal gases; in real gas mixtures, intermolecular interactions make this untrue.
  • Practical calculation: at low to moderate pressures, the virial equation with second virial coefficients (including mixed coefficients Bᵢⱼ) describes real gas mixture behavior adequately.

🎯 Partial pressure and Dalton's Law

🎯 Definition of partial pressure

Partial pressure pᵢ of substance i in a gas mixture: the product of its mole fraction in the gas phase and the pressure of the phase: pᵢ = yᵢ p.

  • This definition applies to any gas mixture, ideal or real.
  • It is a defined quantity, not a measured pressure.
  • Example: if substance A has mole fraction 0.3 in a mixture at 2 bar total pressure, its partial pressure is 0.6 bar.

⚖️ Dalton's Law

  • The sum of all partial pressures equals the total pressure: Σᵢ pᵢ = p.
  • Derivation: Σᵢ pᵢ = Σᵢ yᵢ p = p Σᵢ yᵢ = p (since mole fractions sum to 1).
  • Valid for any gas mixture, whether ideal or real—it follows directly from the definition of partial pressure, not from gas behavior.

🔍 Partial pressure in ideal vs real gases

  • Ideal gas mixture: pᵢ = nᵢ RT / V (the ideal gas equation applied to constituent i alone).
    • This means the partial pressure equals the pressure the pure substance would have at the same T and V if all other constituents were removed.
  • Real gas mixture: partial pressure is still defined as yᵢ p, but it does not equal the pressure of the pure substance at the same T and V, because intermolecular interactions differ between the mixture and the pure gas.
  • Don't confuse: the "pressure if alone" interpretation is only valid for ideal gases.

🧪 Ideal gas mixtures

🧪 Thermodynamic definition

An ideal gas mixture is defined as any gas mixture in which each constituent i obeys μᵢ = μᵢ°(g) + RT ln(pᵢ/p°) at all compositions.

  • μᵢ°(g) is the standard chemical potential: the chemical potential of pure gaseous i at standard pressure p° (1 bar), behaving as an ideal gas, at the same temperature as the mixture.
  • This equation is taken as the definition of an ideal gas mixture, not derived from first principles.
  • The excerpt justifies it with a thought experiment involving a semipermeable membrane separating pure A from a mixture of A and B, both ideal gases, showing that A's partial pressure in the mixture equals the pressure of pure A when they are in equilibrium.

🔬 Physical meaning of ideal behavior

  • An ideal gas has negligible intermolecular interactions.
  • Because molecules don't interact, adding substance B to pure A does not change A's behavior—A "doesn't notice" B is there.
  • The definition via the chemical potential equation is consistent with the ideal gas equation pV = nRT and with internal energy depending only on temperature.

📐 Partial molar quantities in ideal gas mixtures

All partial molar quantities have simple forms:

QuantityFormulaKey feature
Chemical potential μᵢμᵢ°(g) + RT ln(pᵢ/p°)Depends on partial pressure
Partial molar entropy SᵢSᵢ° - R ln(pᵢ/p°)Decreases with increasing partial pressure
Partial molar enthalpy HᵢHᵢ°Independent of pressure and composition
Partial molar volume VᵢRT/pDepends on total pressure, not composition
Partial molar internal energy UᵢUᵢ°Independent of pressure and composition
Partial molar heat capacity Cp,ᵢCp,ᵢ°Independent of pressure and composition
  • The key insight: Hᵢ, Uᵢ, and Cp,ᵢ are functions only of temperature in an ideal gas mixture.
  • Vᵢ = RT/p is the same for all constituents at given T and p (but not necessarily equal to the standard molar volume Vᵢ° = RT/p°).

🎓 Standard state clarification

  • The standard state of substance i in a gas mixture is pure gaseous i at p°, behaving as an ideal gas, at the same temperature as the mixture.
  • If a constituent in an ideal gas mixture has partial pressure equal to p°, its chemical potential equals the standard chemical potential, but it is not in its standard state (because the standard state is the pure gas, not a mixture component).
  • Don't confuse: "same chemical potential as standard state" ≠ "in the standard state."

🌡️ Real gas mixtures

🌡️ Fugacity concept

Fugacity fᵢ of substance i in a real gas mixture: defined by μᵢ = μᵢ°(g) + RT ln(fᵢ/p°), or equivalently fᵢ = p° exp[(μᵢ - μᵢ°(g))/RT].

  • Fugacity is a "kind of effective partial pressure."
  • It is the partial pressure substance i would have in an ideal gas mixture at the same temperature in which μᵢ has the same value as in the real gas mixture.
  • As pressure approaches zero, fᵢ approaches pᵢ (real gases behave ideally at low pressure).

🔧 Fugacity coefficient

Fugacity coefficient φᵢ: defined by fᵢ = φᵢ pᵢ.

  • The fugacity coefficient measures the deviation from ideal behavior.
  • For an ideal gas, φᵢ = 1 (fugacity equals partial pressure).
  • As pressure approaches zero, φᵢ approaches 1.
  • Calculation formula: ln φᵢ(p') = ∫₀^p' [(Vᵢ/RT) - (1/p)] dp at constant T and composition.
    • Vᵢ is the partial molar volume of i in the mixture.
    • The integral vanishes as p' → 0.

📊 Virial equation for real gas mixtures

At low to moderate pressures, the equation of state is:

  • V/n = RT/p + B
  • Or equivalently: Z = pV/(nRT) = 1 + Bp/RT

Where:

  • n is the total amount of gas (sum of all nᵢ).
  • B is the second virial coefficient, which depends on temperature and composition.

🧬 Composition dependence of virial coefficients

For a binary mixture of A and B:

  • B = yₐ² Bₐₐ + 2yₐ yᵦ Bₐᵦ + yᵦ² Bᵦᵦ

For a general mixture:

  • B = Σᵢ Σⱼ yᵢ yⱼ Bᵢⱼ (where Bᵢⱼ = Bⱼᵢ)

Where:

  • Bᵢᵢ is the second virial coefficient of pure i.
  • Bᵢⱼ (i ≠ j) is a mixed second virial coefficient describing interactions between molecules of i and j.
  • All virial coefficients are functions of temperature only.

🔬 Partial molar volume in real gases

Using the virial equation:

  • Vᵢ = RT/p + B'ᵢ

Where B'ᵢ is defined by:

  • B'ᵢ = 2 Σⱼ yⱼ Bᵢⱼ - B

For a binary mixture:

  • B'ₐ = Bₐₐ + (Bₐₐ + 2Bₐᵦ - Bᵦᵦ) yᵦ²
  • B'ᵦ = Bᵦᵦ + (Bₐₐ + 2Bₐᵦ - Bᵦᵦ) yₐ²

📈 Calculating other partial molar quantities

For a real gas mixture obeying V/n = RT/p + B:

  • ln φᵢ = B'ᵢ p / (RT)
  • μᵢ(p') = μᵢ°(g) + RT ln(p'ᵢ/p°) + ∫₀^p' [Vᵢ - RT/p] dp
  • Sᵢ - Sᵢ° = -R ln(pᵢ/p°) - p dB'ᵢ/dT
  • Hᵢ - Hᵢ° = p [B'ᵢ - T dB'ᵢ/dT]
  • Uᵢ - Uᵢ° = pT dB'ᵢ/dT
  • Cp,ᵢ - Cp,ᵢ° = -pT d²B'ᵢ/dT²

Key insight: unlike ideal gas mixtures, these partial molar quantities do depend on pressure and composition (through B'ᵢ and its temperature derivatives).

🔄 Equilibrium conditions in multicomponent systems

🔄 General equilibrium criteria

In an equilibrium state of a multiphase, multicomponent system without internal partitions:

  • Temperature is uniform throughout the system.
  • Pressure is uniform throughout the system.
  • Each species has a uniform chemical potential in all phases where it is present.

Exception: if a species i' is effectively excluded from a particular phase (e.g., sucrose cannot enter ice crystals, or a nonpolar substance is insoluble in water), there is no equilibrium condition involving the chemical potential of i' in that phase.

🧮 Gibbs–Duhem relations for mixtures

At constant T and p:

  • Σᵢ nᵢ dμᵢ = 0
  • Or equivalently: Σᵢ xᵢ dμᵢ = 0

Meaning: the chemical potentials of different species cannot be varied independently at constant T and p—they are constrained by this relation.

More general form (without constant T and p restriction):

  • -S dT + V dp + Σᵢ nᵢ dμᵢ = 0

🔗 Relations among partial molar quantities

Key equations connecting partial molar quantities:

  • μᵢ = Hᵢ - T Sᵢ (analogous to μ = Hₘ - T Sₘ for pure substances)
  • (∂μᵢ/∂T) at constant p and composition = -Sᵢ
  • (∂μᵢ/∂p) at constant T and composition = Vᵢ
  • Uᵢ = Hᵢ - p Vᵢ
  • Cp,ᵢ = (∂Hᵢ/∂T) at constant p and composition

These are the mixture analogues of relations for pure substances.

⚡ Chemical potential of ions (special case)

  • The chemical potential of an individual ion (e.g., K⁺ or Cl⁻) is defined as the rate at which Gibbs energy changes with the amount of that ion at constant T, p, and amounts of other species.
  • This is a hypothetical process in which the net charge of the mixture would increase—not physically realizable in the laboratory.
  • The chemical potential of an ion is therefore a valid but purely theoretical concept.
  • Example: for a KCl solution, we can write dG = -S dT + V dp + μ(H₂O) dn(H₂O) + μ(K⁺) dn(K⁺) + μ(Cl⁻) dn(Cl⁻), but in practice we can only vary n(H₂O) and n(KCl) independently, not n(K⁺) and n(Cl⁻) independently.
  • The equilibrium condition that an ion has uniform chemical potential across phases applies to situations like Donnan membrane equilibrium, where transfer equilibrium of ions exists between solutions separated by a semipermeable membrane.
55

Liquid and Solid Mixtures of Nonelectrolytes

9.4 Liquid and Solid Mixtures of Nonelectrolytes

🧭 Overview

🧠 One-sentence thesis

Liquid and solid mixtures of nonelectrolytes exhibit predictable behavior through Raoult's law (for solvents and ideal mixtures) and Henry's law (for dilute solutes), with ideal-dilute solutions representing the most commonly useful limiting case where solutes obey Henry's law and solvents obey Raoult's law.

📌 Key points (3–5)

  • Raoult's law: the fugacity (or partial pressure) of a constituent is proportional to its mole fraction, with the proportionality constant being the pure-substance fugacity at the same T and p.
  • Henry's law: in dilute solutions, solute fugacity becomes proportional to its mole fraction (or concentration or molality) as the solute approaches infinite dilution.
  • Ideal mixtures vs. ideal-dilute solutions: ideal mixtures (rare) obey Raoult's law at all compositions; ideal-dilute solutions (common) have solutes obeying Henry's law and solvent obeying Raoult's law when sufficiently dilute.
  • Common confusion: a constituent approaches Henry's law as its mole fraction approaches zero, but approaches Raoult's law as its mole fraction approaches unity—the same substance can follow different laws depending on concentration.
  • Reference states: each type of solution behavior defines a reference state (a hypothetical state at the same T and p) used to express chemical potential and calculate partial molar properties.

🧪 Raoult's law and ideal mixtures

🔬 Raoult's law for partial pressure

Raoult's law (original, 1888): when a dilute liquid solution of a volatile solvent and a nonelectrolyte solute is equilibrated with a gas phase, the partial pressure p_A of the solvent in the gas phase is proportional to the mole fraction x_A of the solvent in the solution: p_A = x_A · p_A, where p_A is the saturation vapor pressure of the pure solvent.

  • The modern rigorous form: p_i = x_i · p_i, where p_i is the partial pressure of pure i at the same T and p as the mixture.
  • This applies when you adjust total pressure (e.g., with an inert gas) to keep conditions constant.

🌫️ Raoult's law for fugacity

  • A more general form replaces partial pressures with fugacities: f_i = x_i · f*_i.
  • Here f*_i is the fugacity of pure liquid i at the same T and p as the mixture.
  • The two forms are equivalent when the gas phase is an ideal gas mixture.
  • This fugacity form can be rewritten in terms of chemical potential: μ_i = μ*_i + RT ln x_i, where μ*_i is the chemical potential of pure liquid i.

🎯 Ideal mixtures defined

Ideal liquid mixture: a liquid mixture in which, at a given temperature and pressure, each constituent obeys Raoult's law for fugacity over the entire range of composition.

  • Requirements: constituents must have similar molecular size and structure, and pure liquids must be miscible in all proportions.
  • Example: benzene and toluene form nearly ideal mixtures.
  • Counter-example: water and methanol are miscible but deviate considerably from Raoult's law.
  • Most organic liquid mixtures show positive deviation (higher fugacity than Raoult's law predicts).
  • The same definition applies to ideal solid mixtures (e.g., some metal alloys, gemstones, doped semiconductors).

🧮 Partial molar quantities in ideal mixtures

For any constituent i in an ideal mixture:

PropertyFormulaMeaning
Chemical potentialμ_i = μ*_i + RT ln x_iVaries with composition
EntropyS_i = S*_i - R ln x_iGreater than pure substance (ln x_i is negative)
EnthalpyH_i = H*_iIndependent of composition
VolumeV_i = V*_iIndependent of composition
Internal energyU_i = U*_iIndependent of composition
Heat capacityC_p,i = C*_p,iIndependent of composition
  • At constant T and p, only chemical potential and entropy vary with composition in an ideal mixture.
  • All other partial molar quantities equal the pure-substance values.

💧 Henry's law and dilute solutions

📉 Henry's law behavior

Henry's law: as the mole fraction x_i of a constituent in a liquid mixture approaches zero (at constant T and p), its fugacity f_i in an equilibrated gas phase becomes proportional to x_i: f_i → k_H,i · x_i as x_i → 0.

  • k_H,i is the Henry's law constant, which depends on temperature, total pressure, and the identity of other constituents.
  • This applies to volatile nonelectrolytes; electrolytes behave very differently.
  • The limiting slope of f_i vs. x_i is finite (not zero or infinite).

📊 Three versions of Henry's law

Because mole fraction, concentration, and molality become proportional in the limit of infinite dilution, Henry's law has three equivalent forms:

BasisFormulaHenry's constant definitionStandard composition
Mole fractionf_B = k_H,B · x_Bk_H,B = lim (f_B / x_B) as x_B → 0x° = 1
Concentrationf_B = k_c,B · c_Bk_c,B = lim (f_B / c_B) as c_B → 0c° = 1 mol/dm³
Molalityf_B = k_m,B · m_Bk_m,B = lim (f_B / m_B) as m_B → 0m° = 1 mol/kg
  • The constants are related: k_c,B = V_A · k_H,B and k_m,B = M_A · k_H,B (where V_A is molar volume of pure solvent, M_A is molar mass of solvent).
  • To evaluate a Henry's constant: plot f_B divided by the composition variable vs. the composition variable, then extrapolate to infinite dilution.

🧪 Ideal-dilute solutions

Ideal-dilute solution: a real solution that is dilute enough for each solute to obey Henry's law.

  • Microscopic requirement: solute molecules must be sufficiently separated that solute–solute interactions are negligible.
  • Not the same as an ideal mixture: few mixtures are ideal, but any nonelectrolyte solution becomes ideal-dilute when sufficiently dilute.
  • In this range, solute chemical potential: μ_B = μ_ref,x,B + RT ln x_B (or similar forms for concentration/molality basis).
  • μ_ref,x,B is the chemical potential in a reference state—a hypothetical solution at x_B = 1 that still behaves as ideal-dilute (sometimes called "ideal-dilute solution of unit solute mole fraction").

🔄 Solvent behavior in ideal-dilute solutions

Using the Gibbs–Duhem equation, one can show that in an ideal-dilute solution of nonelectrolytes, each solute obeys Henry's law and the solvent obeys Raoult's law.

  • Equivalent statement: a nonelectrolyte constituent approaches Henry's law as its mole fraction approaches zero, and approaches Raoult's law as its mole fraction approaches unity.
  • Example: ethanol in ethanol-water mixtures shows positive deviations from Raoult's law and negative deviations from Henry's law at intermediate compositions.
  • Don't confuse: the same substance follows different laws depending on whether it is dilute (Henry's law) or concentrated (Raoult's law).

📐 Partial molar quantities in ideal-dilute solutions

For the solvent A (in the ideal-dilute range):

  • Same formulas as for an ideal mixture.
  • Chemical potential and entropy vary with composition; all other properties equal pure-solvent values.

For the solute B:

PropertyFormulaBehavior
Chemical potentialμ_B = μ_ref,x,B + RT ln x_BVaries with composition; → -∞ as x_B → 0
EntropyS_B = S_ref,x,B - R ln x_BVaries with composition; → +∞ as x_B → 0
EnthalpyH_B = H^∞_BConstant (equals value at infinite dilution)
VolumeV_B = V^∞_BConstant (equals value at infinite dilution)
Internal energyU_B = U^∞_BConstant (equals value at infinite dilution)
Heat capacityC_p,B = C^∞_p,BConstant (equals value at infinite dilution)
  • At constant T and p in the ideal-dilute range, only chemical potential and entropy of the solute vary with composition.
  • When pressure equals standard pressure p°, the infinite-dilution values become standard values (H°_B, V°_B, etc.).

🎯 Reference states and activity coefficients

🏷️ Reference states for nonelectrolytes

Reference state: a state of a constituent that has the same temperature and pressure as the mixture; the chemical potential in this state depends only on T and p, not on composition.

  • When pressure equals standard pressure p°, the reference state becomes the standard state (chemical potential = standard chemical potential μ°_i, function of T only).
  • Reference states are useful for derivations at constant T and p when pressure is not necessarily standard.

Types of reference states:

Constituent typeReference state descriptionChemical potential symbol
Substance i in gas mixturePure i as ideal gasμ_ref,i(g)
Substance i in liquid/solid mixturePure i in same physical stateμ*_i
Solvent APure A in same physical stateμ*_A
Solute B (mole fraction)B at x_B = 1, ideal-dilute behaviorμ_ref,x,B
Solute B (concentration)B at c° = 1 mol/dm³, ideal-dilute behaviorμ_ref,c,B
Solute B (molality)B at m° = 1 mol/kg, ideal-dilute behaviorμ_ref,m,B

🔧 Activity coefficients concept

  • An activity coefficient is an adjustment factor relating actual behavior to ideal behavior at the same T and p.
  • The ideal behavior is based on the reference state for the species.
  • Example: in the reference state for a solute on mole fraction basis, the hypothetical solution continues ideal-dilute behavior all the way to x_B = 1, with fugacity f_B = k_H,B (the Henry's law constant).
  • Don't confuse: the reference state is often fictitious (cannot actually be prepared), but it provides a well-defined thermodynamic baseline for calculations.
56

Activity Coefficients in Mixtures of Nonelectrolytes

9.5 Activity Coefficients in Mixtures of Nonelectrolytes

🧭 Overview

🧠 One-sentence thesis

Activity coefficients serve as adjustment factors that relate the actual chemical behavior of mixture components to their ideal behavior at the same temperature and pressure, with their values depending on the choice of reference state and approaching unity under conditions where the mixture behaves ideally.

📌 Key points (3–5)

  • What activity coefficients do: they multiply composition variables (mole fraction, concentration, molality) to give "effective" values that account for non-ideal behavior.
  • Reference states define the baseline: each component has a reference state at the same T and p as the mixture; when p equals the standard pressure, the reference state becomes the standard state.
  • Different reference states for different roles: pure substances for solvents and mixture components; hypothetical extrapolated states (at infinite dilution) for solutes.
  • Common confusion—solute reference state basis: the value of the chemical potential μ_B is the same regardless of basis (mole fraction, concentration, molality), but the activity coefficient value differs depending on which reference state you choose.
  • Limiting behavior: activity coefficients approach 1 under ideal conditions—pure component limit for solvents/mixture constituents, infinite dilution limit for solutes.

🎯 Reference states and standard states

🎯 What a reference state is

A reference state of a constituent has the same temperature and pressure as the mixture; when the species is in its reference state, its chemical potential depends only on T and p of the mixture.

  • The reference state is the baseline for comparing actual behavior to ideal behavior.
  • When pressure equals the standard pressure p°, the reference state becomes the standard state, where the chemical potential is the standard chemical potential μ°_i (function of T only).
  • Reference states are useful for processes at constant T and p when p is not necessarily the standard pressure.

🧪 Reference states for different types of components

The excerpt provides a table of reference states for nonelectrolytes:

Component typeReference stateChemical potential symbol
Substance i in a gas mixturePure i behaving as an ideal gasμ_ref,i(g)
Substance i in liquid/solid mixturePure i in same physical state as mixtureμ*_i
Solvent A of a solutionPure A in same physical state as solutionμ*_A
Solute B, mole fraction basisB at mole fraction 1, behavior extrapolated from infinite dilution (hypothetical)μ_ref,x,B
Solute B, concentration basisB at concentration c°, extrapolated from infinite dilution (hypothetical)μ_ref,c,B
Solute B, molality basisB at molality m°, extrapolated from infinite dilution (hypothetical)μ_ref,m,B
  • Note: solute reference states on concentration or molality basis are hypothetical—they extrapolate ideal-dilute behavior to a standard composition.
  • The subscripts x, c, or m indicate the basis of the reference state.

🧩 Ideal behavior as the starting point

🧩 General form for ideal mixtures

For condensed phases, the chemical potential in an ideal mixture has the form:

μ^id_i = μ_ref,i + RT ln(composition variable / standard composition)

  • For mole fraction basis, standard composition is x° = 1.
  • For concentration basis, standard composition is c°.
  • For molality basis, standard composition is m°.
  • The excerpt lists six equations (9.5.4–9.5.9) for different types of components in ideal mixtures and ideal-dilute solutions.

🔍 Examples of ideal expressions

  • Constituent of ideal gas mixture: μ^id_i(g) = μ_ref,i(g) + RT ln(p_i / p)
  • Constituent of ideal liquid/solid mixture: μ^id_i = μ*_i + RT ln(x_i)
  • Solvent of ideal-dilute solution: μ^id_A = μ*_A + RT ln(x_A)
  • Solute, mole fraction basis: μ^id_B = μ_ref,x,B + RT ln(x_B)
  • Solute, concentration basis: μ^id_B = μ_ref,c,B + RT ln(c_B / c°)
  • Solute, molality basis: μ^id_B = μ_ref,m,B + RT ln(m_B / m°)

All follow the same pattern: reference chemical potential + RT times the logarithm of a dimensionless composition ratio.

🔧 Activity coefficients in real mixtures

🔧 How activity coefficients adjust for non-ideality

The activity coefficient is a dimensionless quantity whose value depends on temperature, pressure, mixture composition, and the choice of reference state; it multiplies the composition variable inside the logarithm to give the actual chemical potential.

  • For real mixtures, the general form becomes: μ_i = μ_ref,i + RT ln[(activity coefficient) × (composition variable / standard composition)]
  • When the mixture behaves ideally, the activity coefficient equals 1 and the expression reduces to the ideal form.
  • Otherwise, the activity coefficient takes whatever value gives the actual chemical potential.

🧮 Expressions for different components

The excerpt provides equations (9.5.13–9.5.18) for real mixtures:

ComponentActivity coefficient symbolExpression
Gas mixture constituentγ_i (fugacity coefficient)μ_i = μ_ref,i(g) + RT ln(γ_i p_i / p)
Liquid/solid mixture constituentγ_iμ_i = μ*_i + RT ln(γ_i x_i)
Solventγ_Aμ_A = μ*_A + RT ln(γ_A x_A)
Solute, mole fractionγ_x,Bμ_B = μ_ref,x,B + RT ln(γ_x,B x_B)
Solute, concentrationγ_c,Bμ_B = μ_ref,c,B + RT ln(γ_c,B c_B / c°)
Solute, molalityγ_m,Bμ_B = μ_ref,m,B + RT ln(γ_m,B m_B / m°)
  • Different symbols (γ_i, γ_A, γ_x,B, γ_c,B, γ_m,B) are used to avoid confusion and to indicate the reference state basis.
  • The book uses γ_i and γ_A instead of the IUPAC-recommended f_i to avoid confusion with fugacity.

🎭 Interpreting products as "effective" quantities

The excerpt suggests interpreting the products in the logarithms as effective composition variables:

  • γ_i p_i is an effective partial pressure.
  • γ_i x_i, γ_A x_A, γ_x,B x_B are effective mole fractions.
  • γ_c,B c_B is an effective concentration.
  • γ_m,B m_B is an effective molality.

Example: The value of γ_i x_i is the mole fraction that would give the same chemical potential in an ideal mixture as the actual chemical potential in the real mixture.

  • These effective variables are alternative ways to express escaping tendency (related exponentially to chemical potential).

⚠️ Common confusion—same μ_B, different γ values

For a solute B in a given solution:

  • The three expressions (mole fraction, concentration, molality basis) must all give the same value of μ_B (the rate at which Gibbs energy increases with added B at constant T and p).
  • However, the activity coefficient values γ_x,B, γ_c,B, and γ_m,B are different because they are defined relative to different reference states.

Don't confuse: the chemical potential is a physical property with one value; the activity coefficient is a mathematical adjustment factor whose numerical value depends on your choice of basis.

📉 Limiting behavior of activity coefficients

📉 When activity coefficients approach unity

As a mixture approaches ideal behavior, each activity coefficient must approach 1:

ComponentConditionLimit
Gas mixture constituentp → 0γ_i → 1
Liquid/solid mixture constituentx_i → 1 (pure component)γ_i → 1
Solventx_A → 1 (pure solvent)γ_A → 1
Solute, mole fractionx_B → 0 (infinite dilution)γ_x,B → 1
Solute, concentrationc_B → 0 (infinite dilution)γ_c,B → 1
Solute, molalitym_B → 0 (infinite dilution)γ_m,B → 1
  • For solvents and mixture components, the limit is the pure substance.
  • For solutes, the limit is infinite dilution.
  • These limits define the conditions under which the reference state behavior (ideal or ideal-dilute) is approached.

🧬 Behavior in dilute solutions—linear dependence

The excerpt presents a theoretical argument based on molecular interactions:

  • In a dilute solution at constant T and p, the number of solute-solute interactions is proportional to the solute mole fraction x_B.
  • Partial molar quantities U_B and V_B should be approximately linear in x_B.
  • The chemical potential can be written as μ_B = μ_ref,x,B + RT ln(x_B) + k_x x_B, where k_x is a constant related to solute-solute interactions.
  • Solving for the activity coefficient gives: γ_x,B = exp(k_x x_B / RT) ≈ 1 + (k_x / RT) x_B + ...
  • At low x_B, the activity coefficient is a linear function of x_B.

Similarly, for molality basis:

  • γ_m,B = 1 + (k_m / RT) m_B + ... at low m_B.

Example: An ideal-dilute solution is one where x_B is much smaller than RT / k_x, so γ_x,B ≈ 1.

  • The excerpt notes this prediction is confirmed experimentally (Figure 9.10).
  • Important distinction: this linear behavior applies only to nonelectrolyte solutes; for electrolytes, the slope approaches −∞ at low molality (mentioned for later chapter).

🔬 Evaluating activity coefficients from gas fugacities

🔬 Using equilibrium between liquid and gas

When a liquid mixture is equilibrated with a gas phase, if component i is volatile, its fugacity f_i in the gas phase can be used to evaluate the activity coefficient γ_i in the liquid.

  • The chemical potential must be the same in both phases at equilibrium.
  • Equate the expressions for μ_i in the liquid (Eq. 9.5.14) and gas (Eq. 9.5.11), then solve for γ_i.
  • The result is γ_i = C_i (f_i / x_i), where C_i depends on T and p but not on composition.

🧪 Deriving the fugacity-based formula

To find C_i:

  • Use the limiting condition γ_i → 1 as x_i → 1 (pure component limit).
  • At x_i = 1, f_i becomes f*_i (fugacity of pure liquid i at the same T and p).
  • Therefore C_i = 1 / f*_i.
  • Substituting back gives: γ_i = f_i / (x_i f*_i)

For solutes, similar derivations using the infinite-dilution limits (where activity coefficients approach 1) lead to:

  • Solvent: γ_A = f_A / (x_A f*_A)
  • Solute, mole fraction: γ_x,B = f_B / (k_H,B x_B), where k_H,B is the Henry's law constant.
  • Solute, concentration: γ_c,B = f_B / (k_c,B c_B)
  • Solute, molality: γ_m,B = f_B / (k_m,B m_B)

The excerpt provides these in Table 9.4.

📊 Example—ethanol and water

Figure 9.9 illustrates the method for ethanol (A) in ethanol–water mixtures at 25°C:

  • Plot f_A versus x_A.
  • At x_A = 1 (pure ethanol), f_A = f*_A (filled circle, the reference state).
  • At x_A = 0.4 (open circles), measure f_A and compare to x_A f*_A (the Raoult's law prediction).
  • The ratio γ_A = f_A / (x_A f*_A) gives the activity coefficient.
  • As x_A → 1, γ_A → 1, as required.

📊 Example—1-butanol in water

Figure 9.10 shows 1-butanol (B) as a solute in water at 50.08°C, with limited solubility:

  • Plot f_B versus x_B (or m_B).
  • Extrapolate the linear (Henry's law) region to the standard composition to locate the reference state (filled circles).
  • At the mole-fraction reference state (x_B = 1, hypothetical), f_B = k_H,B.
  • At the molality reference state (m_B = m°, hypothetical), f_B = k_m,B m°.
  • The fugacity values at the two reference states are quite different.
  • Activity coefficients γ_x,B and γ_m,B approach 1 at infinite dilution and vary linearly with x_B or m_B in the dilute region, confirming the theoretical prediction.

Don't confuse: the reference state for a solute is not a real state you can prepare; it is a hypothetical extrapolation from infinite dilution.

🔄 Using the Gibbs–Duhem equation

🔄 When one component is non-volatile

If component B of a binary liquid mixture has low volatility, direct fugacity measurement is impractical.

  • If component A is volatile, measure f_A over a range of compositions to get γ_A.
  • Use the Gibbs–Duhem equation to calculate γ_B indirectly.

🧮 The Gibbs–Duhem relation for binary mixtures

At constant T and p:

x_A dμ_A + x_B dμ_B = 0

  • This relates the changes in chemical potentials of the two components during a composition change.
  • Taking the differential of μ_A = μ*_A + RT ln(γ_A x_A) at constant T and p gives dμ_A = RT d ln γ_A + RT d ln x_A.
  • (The excerpt begins this derivation but is cut off; the method involves integrating the Gibbs–Duhem equation to find γ_B from known γ_A data.)

Example scenario: You have fugacity data for volatile component A across all compositions. Calculate γ_A = f_A / (x_A f*_A) at each composition. Use the Gibbs–Duhem equation to integrate and find γ_B, even though you cannot measure f_B directly.

57

Evaluation of Activity Coefficients

9.6 Evaluation of Activity Coefficients

🧭 Overview

🧠 One-sentence thesis

Activity coefficients of mixture components can be evaluated experimentally through fugacity measurements, osmotic coefficients, or the Gibbs–Duhem equation, with the choice of method depending on the volatility and miscibility of the components.

📌 Key points (3–5)

  • Activity coefficients approach unity at their reference conditions: γ → 1 as x → 1 for pure-liquid reference states, and γ → 1 at infinite dilution for solute reference states.
  • Fugacity measurements in equilibrated gas phases provide the most direct route to activity coefficients for volatile components.
  • Gibbs–Duhem equation allows calculation of a nonvolatile component's activity coefficient from measurements on a volatile component in the same mixture.
  • Osmotic coefficient method is especially useful for nonvolatile solutes, relating the solvent chemical potential lowering to the solute activity coefficient through integration.
  • Common confusion: Different reference states (pure liquid vs. solute) yield different numerical values for fugacity and activity coefficients at the same composition; the choice depends on whether components mix in all proportions.

📊 Behavior of activity coefficients

📈 Limiting behavior at reference conditions

Activity coefficient: the ratio of fugacity to the product of composition variable and reference-state fugacity.

  • For a pure-liquid reference state: γ_A = f_A / (x_A · f°_A)
  • As x_A approaches 1 (pure A), γ_A must approach 1 according to the defining equations.
  • For a solute reference state: γ approaches 1 at infinite dilution (x → 0 or m → 0).
  • This limiting behavior is built into the reference-state definitions and serves as a consistency check on experimental data.

🔍 Dilute solution behavior

Figure 9.10 demonstrates 1-butanol (limited solubility in water) at 50.08 °C:

  • Mole fraction basis: Henry's law constant k_H,B defines the reference state; the filled circle shows where f_B equals k_H,B.
  • Molality basis: A different Henry's law constant k_m,B · m° defines the reference state; f_B at the reference state differs from the mole-fraction case.
  • Activity coefficients: Both γ_x,B and γ_m,B approach 1 at infinite dilution and vary linearly with composition in the dilute region, as predicted by theory.
  • Don't confuse: The same solute has different reference fugacities and different activity coefficient values depending on the composition scale chosen.

🔗 Gibbs–Duhem method for nonvolatile components

🧮 Deriving component B from component A

When component B has low volatility but component A is volatile:

  • Measure f_A over a range of liquid compositions to evaluate γ_A = f_A / (x_A · f°_A).
  • The Gibbs–Duhem equation at constant T and p relates changes in chemical potentials: x_A · dμ_A + x_B · dμ_B = 0.
  • Express dμ in terms of activity coefficients: dμ_A = RT · d ln γ_A + (RT / x_A) · dx_A.
  • Rearrange to obtain: d ln γ_B = −(x_A / x_B) · d ln γ_A.

📐 Integration formula

The activity coefficient of B at composition x'_B is given by:

ln γ_B(x'_B) = − ∫[from x_B=1 to x_B=x'_B] (x_A / x_B) · d ln γ_A

  • Integration starts at x_B = 1 (pure B) where γ_B = 1 and ln γ_B = 0.
  • This method works for binary mixtures that mix in all proportions (both components use pure-liquid reference states).
  • Limitation: If B has limited solubility, the integration starting point x_B = 0 makes the integrand x_A / x_B infinite; use the osmotic coefficient method instead.

🌊 Osmotic coefficient method

🧪 Definition and physical meaning

Osmotic coefficient φ_m: a dimensionless function defined by φ_m = (μ°_A − μ_A) / (RT · M_A · Σ m_i), where the sum is over all solute species.

  • The numerator is the lowering of solvent chemical potential caused by solutes.
  • The denominator is the ideal-dilute prediction of this lowering.
  • Deviation from unity: φ_m = 1 at infinite dilution; φ_m ≠ 1 measures the deviation from ideal-dilute behavior.
  • The name "osmotic coefficient" comes from its relation to osmotic pressure: Π / m_B equals φ_m times the limiting value at infinite dilution.

📏 Evaluating φ_m from measurements

Any method that measures μ°_A − μ_A allows evaluation of φ_m:

  • Fugacity measurements: μ°_A − μ_A = RT · ln(f°_A / f_A), so measuring f_A in a gas phase equilibrated with the solution gives φ_m directly.
  • Freezing point depression and osmotic pressure measurements (described in later sections) also yield φ_m.
  • Example: For aqueous sucrose, φ_m has been measured from zero molality up to several mol/kg.

🔢 Calculating solute activity coefficient from φ_m

For a binary nonelectrolyte solution, the solute activity coefficient on a molality basis is:

ln γ_m,B(m'_B) = φ_m(m'_B) − 1 + ∫[from 0 to m'_B] [(φ_m − 1) / m_B] · dm_B

  • Integration starts at m_B = 0 (infinite dilution) where γ_m,B = 1.
  • The integrand (φ_m − 1) / m_B is a slowly varying function that approaches a finite value as m_B → 0 for nonelectrolytes.
  • Figure 9.11 shows for aqueous sucrose:
    • (a) The integrand (φ_m − 1) / m_B varies smoothly from 0 to ~0.1 kg/mol over 0–5 mol/kg.
    • (b) γ_m,B rises above 1 beyond the dilute region, reaching ~2.5 at 5 mol/kg.
    • (c) The effective molality γ_m,B · m_B becomes much larger than actual m_B at high concentration.

🔬 Experimental fugacity measurement techniques

💨 Direct methods for volatile components

Various techniques measure partial pressure p_i in a gas phase equilibrated with a liquid mixture:

  • Manometry: For nonvolatile solutes, evacuate air above the solution and measure total pressure (= solvent partial pressure).
  • Dynamic methods: Pass inert gas through the liquid, then analyze the gas mixture; example: pass dry air through aqueous solution, then through desiccant, and weigh the water absorbed.
  • Gas nonideality correction: Convert measured partial pressure to fugacity using corrections like f_i = φ_i · p_i (φ from equations of state).

⚖️ Isopiestic vapor pressure technique

A comparative method for determining water fugacity in aqueous solutions:

  • Setup: Place dishes containing water + solute B and water + reference solute (known properties, e.g., sucrose, NaCl, CaCl₂) in wells of a metal block for thermal equilibration.
  • Procedure: Evacuate air, seal the chamber, rock gently in a thermostat for days to weeks.
  • Equilibration: Water transfers through vapor until chemical potential of water is the same in all dishes (solutions are "isopiestic").
  • Measurement: Weigh dishes to determine final molality of each solution.
  • Result: Water fugacity is known as a function of reference-solute molality, and this equals the water fugacity over solute B at its measured molality.
  • Works for both nonelectrolyte and electrolyte solutes; can be adapted to nonaqueous solvents.

🎯 Activity definition and standard states

📖 Activity of an uncharged species

Activity a_i: a dimensionless quantity defined by a_i = exp[(μ_i − μ°_i) / (RT)], or equivalently μ_i = μ°_i + RT · ln a_i.

  • μ°_i is the standard chemical potential of species i.
  • Activity depends on temperature, pressure, and composition, and on the choice of standard state.
  • Also called "relative activity" because it is relative to a standard state.
  • Important property: When μ_i = μ°_i (i.e., in the standard state), ln a_i = 0 and therefore a_i = 1.

🏛️ Standard states for different mixture types

Standard states have the same definitions as reference states (Table 9.3) with the additional requirement that pressure equals the standard pressure p°:

Component typeStandard stateNotation
Gas mixturePure ideal gas at T, p°μ°_i(g)
Liquid/solid mixturePure liquid/solid at T, p°μ°_i
Solvent APure liquid A at T, p°μ°_A
Solute B (mole fraction)Hypothetical ideal-dilute at x_B=1, T, p°μ°_x,B
Solute B (concentration)Hypothetical ideal-dilute at c_B=c°, T, p°μ°_c,B
Solute B (molality)Hypothetical ideal-dilute at m_B=m°, T, p°μ°_m,B

🔗 Relating activity to composition

Activity equals the product of three factors:

a_i = Λ_i · γ_i · (composition / standard composition)

  • Pressure factor Λ_i: defined by Λ_i = exp[(μ^ref_i − μ°_i) / (RT)]; equals 1 when p = p°; depends only on pressure at fixed T.
  • Activity coefficient γ_i: measures deviation from ideal behavior at the given composition.
  • Composition ratio: x_i for mixtures, m_B / m° for solute on molality basis, etc.

Example for solute B on molality basis:

a_m,B = Λ_m,B · γ_m,B · (m_B / m°)

Table 9.5 in the excerpt gives explicit expressions for all common cases, including relations to fugacity (e.g., a_m,B = Λ_m,B · f_B / (k_m,B · m°)).

58

Activity of an Uncharged Species

9.7 Activity of an Uncharged Species

🧭 Overview

🧠 One-sentence thesis

The activity of a mixture component is the product of a pressure factor, an activity coefficient, and a composition variable divided by the standard composition, allowing chemical potential to be related to both composition and pressure.

📌 Key points

  • What activity relates: Activity connects the chemical potential of a component to its composition and pressure through a standard state.
  • Pressure factor role: The pressure factor (Γ) accounts for the difference between the reference state and standard state pressures; it equals 1 at standard pressure and differs significantly from 1 only at high pressures.
  • General activity formula: For condensed-phase mixtures, activity = (pressure factor) × (activity coefficient) × (composition / standard composition).
  • Common confusion: Reference state vs. standard state—the reference state is at the mixture's actual pressure, while the standard state is at standard pressure p°; the pressure factor bridges them.
  • Practical simplification: At moderate pressures (near 1 bar), the pressure factor is close to unity and is often omitted from activity expressions.

🔗 Chemical potential and activity relationships

🔗 Basic activity definitions

The excerpt provides chemical potential expressions for different types of components:

Component typeChemical potential formula
Gas mixture component iμ_i(g) = μ°_i(g) + RT ln a_i(g)
Liquid/solid mixture component iμ_i = μ°_i + RT ln a_i
Solvent Aμ_A = μ°_A + RT ln a_A
Solute B (mole fraction)μ_B = μ°_x,B + RT ln a_x,B
Solute B (concentration)μ_B = μ°_c,B + RT ln a_c,B
Solute B (molality)μ_B = μ°_m,B + RT ln a_m,B
  • Each formula relates the chemical potential at actual conditions to the standard-state chemical potential through the activity.
  • The standard state chemical potential (μ°) depends on the basis chosen (mole fraction, concentration, or molality for solutes).

🔗 Why different bases matter

  • Solutes can be described using mole fraction (x_B), concentration (c_B), or molality (m_B).
  • Each basis has its own standard state and therefore its own standard chemical potential and activity expression.
  • The choice depends on convenience: molality is temperature-independent, concentration is common in kinetics, mole fraction is symmetric for all components.

🔢 The pressure factor concept

🔢 Definition and meaning

Pressure factor Γ_i: a dimensionless quantity defined as Γ_i = exp[(μ^ref_i - μ°_i) / RT], where μ^ref_i is the chemical potential in the reference state and μ°_i is the standard-state chemical potential.

  • The excerpt calls Γ_i the "pressure factor" (Pitzer and Brewer called it "the activity in a reference state").
  • At a given temperature, Γ_i depends only on the pressure p of the mixture.
  • Γ_i equals 1 when the mixture pressure p equals the standard pressure p°.
  • The pressure factor accounts for the fact that the reference state is at the mixture's actual pressure, while the standard state is at p°.

🔢 Why it exists

  • The reference state and standard state are both at the same temperature but may be at different pressures.
  • If the mixture is not at standard pressure, the chemical potentials of these two states differ.
  • The pressure factor quantifies this difference and allows activity to be expressed in terms of composition variables.
  • Don't confuse: the pressure factor is not about how pressure affects composition; it's about how pressure affects the chemical potential difference between reference and standard states.

📐 Activity and composition relationship

📐 Deriving the activity formula

The excerpt uses solute B on a molality basis as an example:

  • Starting from μ_B = μ^ref_m,B + RT ln(γ_m,B m_B / m°) and μ_B = μ°_m,B + RT ln a_m,B
  • Equating these gives: ln a_m,B = (μ^ref_m,B - μ°_m,B) / RT + ln(γ_m,B m_B / m°)
  • Recognizing (μ^ref_m,B - μ°_m,B) / RT = ln Γ_m,B
  • Therefore: a_m,B = Γ_m,B γ_m,B (m_B / m°)

📐 General pattern for activities

The excerpt states the general rule:

The activity of a constituent of a condensed-phase mixture is in general equal to the product of the pressure factor, the activity coefficient, and the composition variable divided by the standard composition.

Table of explicit activity expressions for nonelectrolytes:

SubstanceActivity expressionAlternative form
Pure gasa(g) = Γ(g) φ = f / p°
Pure liquid or solida = Γ
Gas mixture component ia_i(g) = Γ_i(g) φ_i (p_i / p)= f_i / p°
Liquid/solid mixture component ia_i = Γ_i γ_i x_i= Γ_i (f_i / f*_i)
Solvent Aa_A = Γ_A γ_A x_A= Γ_A (f_A / f*_A)
Solute B (mole fraction)a_x,B = Γ_x,B γ_x,B x_B= Γ_x,B (f_B / k_H,B)
Solute B (concentration)a_c,B = Γ_c,B γ_c,B (c_B / c°)= Γ_c,B (f_B / k_c,B c°)
Solute B (molality)a_m,B = Γ_m,B γ_m,B (m_B / m°)= Γ_m,B (f_B / k_m,B m°)
  • The alternative forms show activities can also be expressed using fugacities (f).
  • For condensed phases, the fugacity refers to the fugacity in a gas phase equilibrated with the condensed phase.

🔬 Pressure dependence of pressure factors

🔬 Deriving pressure factor formulas

To find how Γ_i depends on pressure, the excerpt uses:

  • The defining equation Γ_i = exp[(μ^ref_i - μ°_i) / RT]
  • The relation (∂μ_i / ∂p)_T,{n_i} = V_i (partial molar volume)
  • Integration: μ^ref_i(p') - μ°_i = ∫[p° to p'] V_i dp

For different component types:

Gas mixture component:

  • Reference state: pure gas i at mixture pressure, behaving as ideal gas
  • μ^ref_i(g) = μ°_i(g) + RT ln(p / p°)
  • Therefore: Γ_i(g) = p / p°

Condensed-phase component (pure, solvent, or solute on x or m basis):

  • μ^ref_i(p') - μ°_i = ∫[p° to p'] V_i dp
  • V_i is the molar volume V_i (pure substance), V_A (pure solvent), or V^∞_B (solute at infinite dilution)
  • Γ_i = exp[∫(p° to p') (V_i / RT) dp]
  • Approximate (incompressible): Γ_i ≈ exp[V_i(p' - p°) / RT]

Solute on concentration basis:

  • Requires special treatment because concentration changes slightly with pressure at constant composition (constant mole fraction or molality)
  • Derivation uses isothermal compressibility κ_T = -(1/V)(∂V/∂p)_T,{n_i}
  • Result: μ^ref_c,B(p') - μ°_c,B = ∫[p° to p'] (V^∞_B - RT κ^∞_T) dp
  • Γ_c,B = exp[∫(p° to p') (V^∞_B / RT - κ^∞_T) dp]
  • Approximate: Γ_c,B ≈ exp[V^∞_B(p' - p°) / RT]

🔬 Numerical magnitude

The excerpt provides concrete numbers assuming V*_i = 100 cm³/mol, T = 300 K, p° = 1 bar:

PressureΓ_i value
0 bar (limit)0.996
1 bar1.000
2 bar1.004
10 bar1.04
100 bar1.49
  • At moderate pressures near 1 bar, Γ_i differs negligibly from unity.
  • Only at high pressures (tens to hundreds of bars) does the pressure factor differ appreciably from 1.
  • This is why simplified activity expressions (omitting Γ) are common: a_i = γ_i x_i, a_m,B = γ_m,B (m_B / m°), etc.

🔬 Practical implications

  • For most laboratory work at atmospheric pressure, the pressure factor can be safely ignored.
  • At high pressures, including Γ is essential for accurate thermodynamic calculations.
  • The excerpt notes that choosing a standard pressure p° within the experimental pressure range (e.g., p° = 1 kbar for high-pressure work) keeps Γ close to unity.
  • Don't confuse: the pressure factor corrects for the pressure difference between reference and standard states; it does not describe how composition changes with pressure.

🌍 Mixtures in gravitational fields (brief overview)

🌍 Equilibrium conditions

The excerpt briefly discusses gas mixtures in tall columns under gravity:

  • The system is divided into thin horizontal slab-shaped phases at different elevations.
  • At equilibrium, temperature and the chemical potential of each constituent are uniform throughout the gas mixture.
  • Pressure is not uniform; it varies with elevation.
  • This parallels the treatment of pure gases in gravitational fields.

🌍 Chemical potential variation with elevation

  • The chemical potential of component i at elevation h is the partial molar Gibbs energy at that elevation.
  • At constant T, p, and composition: dμ_i = M_i g dh, where M_i is molar mass and g is gravitational acceleration.
  • Integrating from reference elevation h = 0 to elevation h': μ_i(h') - μ_i(0) = M_i g h'
  • The fugacity f_i remains constant during reversible isothermal elevation at constant volume.
  • This shows how chemical potential (and hence activity) must be defined carefully when gravitational or centrifugal fields are present.
59

Mixtures in Gravitational and Centrifugal Fields

9.8 Mixtures in Gravitational and Centrifugal Fields

🧭 Overview

🧠 One-sentence thesis

At equilibrium, gas mixtures in gravitational fields and liquid solutions in centrifuge cells maintain uniform temperature and chemical potentials throughout while exhibiting spatial variation in pressure and composition due to the applied field.

📌 Key points (3–5)

  • Uniform at equilibrium: Temperature and chemical potential of each component are uniform throughout the system, but pressure and composition vary with position.
  • Gas mixture behavior: Each constituent in an ideal gas mixture obeys the barometric formula individually, causing composition to vary with elevation when components have different molar masses.
  • Centrifuge equilibrium: In a spinning centrifuge, solute concentration varies radially depending on whether the solution density is greater or less than the effective solute density.
  • Common confusion: Chemical potential stays uniform at equilibrium even though concentration changes with position—the position-dependent energy term (gravitational or centrifugal) compensates for the concentration gradient.
  • Practical application: Sedimentation equilibrium in ultracentrifuges allows determination of macromolecule molar mass from measured concentration gradients.

🌍 Gas mixtures under gravity

🔬 Equilibrium conditions

When a tall gas column reaches equilibrium in a gravitational field:

  • Temperature is uniform throughout all elevations
  • Chemical potential of each component is uniform throughout
  • Pressure is not uniform—it varies with height

Chemical potential of substance i in a mixture at elevation h: the partial molar Gibbs energy at that elevation.

The standard potential is defined at a reference elevation h = 0.

📐 How fugacity varies with height

When a gas sample is raised by distance dh at constant temperature and volume:

  • The fugacity f_i remains constant during the reversible elevation process
  • Gravitational work mg dh contributes to internal energy change
  • Integration shows: chemical potential at height h' minus chemical potential at h = 0 equals M_i g h' (where M_i is molar mass, g is gravitational acceleration)

At equilibrium, chemical potential must be equal at all heights, which requires:

  • Fugacity decreases as elevation increases
  • The relationship is: f_i(h) = f_i(0) exp(−M_i g h / RT)

Example: A gas mixture at sea level has higher fugacity for each component than the same mixture at higher elevation.

🎈 Ideal gas mixture behavior

For an ideal gas mixture, fugacity equals partial pressure:

  • p_i(h) = p_i(0) exp(−M_i g h / RT)
  • Each constituent individually obeys the barometric formula
  • Total pressure at height h is the sum of all partial pressures

🔄 Composition changes with elevation

When constituents have different molar masses:

  • The component with greater molar mass has its mole fraction decrease with increasing elevation
  • The component with smaller molar mass has its mole fraction increase with increasing elevation

Don't confuse: The total amount of each substance is fixed, but the spatial distribution changes—heavier molecules concentrate at lower elevations.

🌀 Liquid solutions in centrifuge cells

🎯 Reference frame considerations

The analysis uses a local frame fixed in the spinning rotor:

  • The rotor's angle relative to the lab is not relevant to the system state
  • Equilibrium exists only relative to the rotating frame (an observer in this frame sees no change over time)
  • No equilibrium exists relative to the lab frame (the system's position constantly changes)
  • Thermodynamic laws apply in the rotating frame when measurements are made in this frame

⚖️ Forces in the rotating frame

A body of mass m at radial distance r experiences:

  • Centrifugal force: m ω² r directed outward (where ω is angular velocity)
  • Gravitational force: directed downward (same as in lab frame)

For typical centrifuge cells (height ≤ 1 cm):

  • Variation between top and bottom at any given radius is negligible
  • All measurable variation is along the radial direction

🧪 Equilibrium conditions in the centrifuge

Same conclusion as for gravitational field:

  • Temperature is uniform throughout the solution
  • Chemical potential of each substance (solvent and solute) is uniform throughout
  • Pressure and concentration vary with radial position

📊 Pressure variation with radius

Considering thin slab-shaped volume elements perpendicular to the radial direction:

  • Forces at equilibrium: inner face pressure, outer face pressure, and centrifugal force sum to zero
  • Result: dp = ω² r ρ dr (where ρ is solution density)
  • Integrating: p'' − p' = (ω² ρ / 2)[(r'')² − (r')²]
  • Pressure increases with increasing distance from the rotation axis

🧬 Solute concentration variation

The key relationship for solute B at equilibrium:

ln[c_B(r'') / c_B(r')] = [M_B (1 − v̄_B ρ) ω² / 2RT] [(r'')² − (r')²]

where:

  • M_B is solute molar mass
  • v̄_B is partial specific volume of solute at infinite dilution
  • ρ is solution density
  • The term (1 − v̄_B ρ) acts as a buoyancy factor

🎚️ Direction of concentration gradient

ConditionConcentration behaviorPhysical meaning
ρ < 1/v̄_BConcentration increases with rSolute is effectively denser than solution
ρ > 1/v̄_BConcentration decreases with rSolute is effectively less dense than solution
v̄_B ρ < 1Concentration increases outwardPositive buoyancy factor
v̄_B ρ > 1Concentration increases inwardNegative buoyancy factor

Don't confuse: The buoyancy factor depends on the partial specific volume of the solute, not just the solution density—it's the relative density that matters.

🔬 Sedimentation equilibrium method

📏 Determining molar mass

Sedimentation equilibrium: a method of determining the molar mass of a macromolecule by measuring concentration gradients in a centrifuge at equilibrium.

Procedure:

  1. Place dilute macromolecule solution in an analytical ultracentrifuge cell
  2. Select angular velocity to produce measurable concentration gradient at equilibrium
  3. Measure solute concentration optically as a function of radial position r
  4. Plot ln(c_B / c°) versus r²

Analysis:

  • The plot is predicted to be linear
  • Slope = M_B (1 − v̄_B ρ) ω² / 2RT
  • Partial specific volume v̄_B is found from separate density measurements
  • Molar mass M_B is then calculated from the slope

Example: For a protein solution where the plot of ln(concentration) vs r² gives a straight line, the slope combined with known density and partial specific volume yields the protein's molar mass.

🎓 Why chemical potential stays uniform

The chemical potential expression includes:

  • A concentration-dependent term: RT ln(activity)
  • A position-dependent energy term: gravitational (M_i g h) or centrifugal (−½ M_B ω² [r² − r'²])

At equilibrium, as concentration changes with position, the position-dependent energy term changes in the opposite direction, keeping total chemical potential constant throughout.

60

10.1 Single-Ion Quantities

10.1 Single-Ion Quantities

🧭 Overview

🧠 One-sentence thesis

Sedimentation equilibrium in a centrifuge allows determination of a macromolecule's molar mass by measuring how solute concentration varies with radial distance, with the direction of the gradient depending on whether the solution density is greater or less than the effective solute density.

📌 Key points (3–5)

  • Core relationship: At equilibrium in a centrifuge, the ratio of solute concentrations at two radial positions depends on molar mass, buoyancy, angular velocity, and the difference in radial distances squared.
  • Buoyancy factor: The term (1 − ρv¹ᴮ) acts like a buoyancy factor—it determines whether solute concentration increases or decreases with distance from the rotation axis.
  • Common confusion: The direction of the concentration gradient depends on comparing solution density ρ to the effective solute density 1/v¹ᴮ, not just on centrifugal force alone.
  • Practical application: Sedimentation equilibrium measures molar mass by plotting ln(cᴮ/c°) versus r² to obtain a linear relationship whose slope contains the molar mass.
  • Key assumption: The method assumes the solution is sufficiently dilute that the activity coefficient can be approximated as 1.

🧪 Equilibrium concentration distribution

🧪 The fundamental equation

The excerpt derives the relationship:

ln[cᴮ(r″)/cᴮ(r′)] = Mᴮ(1 − ρv¹ᴮ)ω²/2RT × [(r″)² − (r′)²]

where:

  • cᴮ(r″) and cᴮ(r′) are solute concentrations at radial positions r″ and r′
  • Mᴮ is the molar mass of the solute
  • ρ is the solution density
  • v¹ᴮ is the partial specific volume of the solute at infinite dilution
  • ω is the angular velocity
  • R is the gas constant, T is temperature

How it works:

  • The equation connects activity ratios at two radial positions to the centrifugal field effect.
  • The derivation assumes dilute solution conditions where the activity coefficient γc,B ≈ 1.
  • The pressure factor is incorporated through the relation βc,B ≈ exp[V¹ᴮ(p − p°)/RT].

🔄 From activities to concentrations

The excerpt shows the transformation:

  • Start with activities: ac,B = γc,B × (cᴮ/c°) × βc,B
  • Under dilute conditions: γc,B ≈ 1
  • Substitute the pressure difference from an earlier equation (9.8.12)
  • Use the relation V¹ᴮ = Mᴮv¹ᴮ to connect molar volume to partial specific volume
  • Equate two expressions for ln[ac,B(r″)/ac,B(r′)] to obtain the final concentration equation

⚖️ The buoyancy factor

⚖️ What (1 − ρv¹ᴮ) means

The excerpt describes this factor as "like a buoyancy factor for the effect of the centrifugal field on the solute."

Physical interpretation:

  • Compares solution density ρ to effective solute density 1/v¹ᴮ
  • Determines whether the solute is effectively "heavier" or "lighter" than the surrounding solution in the centrifugal field

🔀 Direction of concentration gradient

The buoyancy factor controls the gradient direction:

ConditionValue of (1 − ρv¹ᴮ)Concentration behavior
ρ < 1/v¹ᴮPositive (< 1)Concentration increases with increasing r
ρ > 1/v¹ᴮNegativeConcentration decreases with increasing r

Don't confuse:

  • The centrifugal force always points outward, but the concentration gradient direction depends on the relative densities.
  • If the solution is denser than the effective solute density, the solute behaves as if it's buoyant and concentrates toward the axis.

📏 Sedimentation equilibrium method

📏 Experimental setup

Sedimentation equilibrium: a method of determining the molar mass of a macromolecule.

Procedure:

  • Place a dilute solution of the macromolecule in an analytical ultracentrifuge cell
  • Select angular velocity ω to produce a measurable solute concentration gradient at equilibrium
  • Measure solute concentration optically as a function of radial position r

📊 Data analysis

Linear plot prediction:

  • The equation predicts that ln(cᴮ/c°) versus r² will be linear
  • Slope = Mᴮ(1 − ρv¹ᴮ)ω²/2RT

Determining molar mass:

  1. Measure the partial specific volume v¹ᴮ from solution density measurements as a function of solute mass fraction (referenced to page 237)
  2. Obtain the slope from the linear plot
  3. Calculate Mᴮ from the slope using known values of ρ, ω, R, T, and v¹ᴮ

Example: An analytical ultracentrifuge spins a dilute protein solution. Optical measurements show how protein concentration varies with distance from the axis. Plotting the natural log of concentration against the square of radial distance yields a straight line. From the slope and independently measured partial specific volume, the protein's molar mass is calculated.

🔬 Key assumptions

  • Dilute solution: Activity coefficient γc,B ≈ 1
  • Equilibrium: The system has reached steady state where concentration distribution no longer changes with time
  • Known parameters: Solution density ρ and partial specific volume v¹ᴮ must be measured independently
61

10.2 Solution of a Symmetrical Electrolyte

10.2 Solution of a Symmetrical Electrolyte

🧭 Overview

🧠 One-sentence thesis

For a symmetrical electrolyte (one cation and one anion per formula unit), the solute chemical potential is the sum of the two ion chemical potentials and depends on the square of the molality, not the first power as in nonelectrolytes, allowing the measurable mean ionic activity coefficient to characterize the solution.

📌 Key points (3–5)

  • What a symmetrical electrolyte is: a strong electrolyte with one cation and one anion per formula unit (ν = 2), such as NaCl (1:1), MgSO₄ (2:2), or HCl.
  • Key difference from nonelectrolytes: the chemical potential depends on the second power of molality (m_B / m°)², not the first power.
  • Mean ionic activity coefficient: the product of the two single-ion activity coefficients (γ_± = √(γ_+ · γ_−)) is measurable, even though individual ion coefficients cannot be determined.
  • Common confusion: individual ion properties (μ_+, μ_−, γ_+, γ_−) cannot be measured experimentally, but their sum or product (for the solute as a whole) can be.
  • Why it matters: the mean ionic activity coefficient can be determined from vapor pressure, osmotic coefficients, or galvanic cell potentials, making thermodynamic analysis of electrolyte solutions practical.

🧪 Defining the symmetrical electrolyte

🧪 What symmetrical means

A symmetrical strong electrolyte: a substance whose formula unit has one cation and one anion that dissociate completely, indicated by ν = 2 (the number of ions per formula unit).

  • Examples from the excerpt:
    • 1:1 salt such as NaCl
    • 2:2 salt such as MgSO₄
    • Strong monoprotic acid such as HCl
  • "Symmetrical" means equal numbers of cations and anions: n_B = n_+ = n_−.
  • The excerpt focuses on binary solutions (solvent A and electrolyte solute B).

🔗 Relating solute and ion chemical potentials

  • The Gibbs energy can be written two ways:

    • G = n_A μ_A + n_B μ_B (treating the solute as a whole)
    • G = n_A μ_A + n_+ μ_+ + n_− μ_− (treating ions separately)
  • Comparing these for a symmetrical electrolyte (n_B = n_+ = n_−) gives:

    μ_B = μ_+ + μ_− (for ν = 2)

  • The solute chemical potential is the sum of the single-ion chemical potentials.

  • Importantly, μ_B does not depend on the electric potential Φ, unlike μ_+ and μ_− individually.

🔢 The squared-molality relationship

🔢 Why the second power appears

  • Substituting the ion expressions (from Eq. 10.1.10) into μ_B = μ_+ + μ_− and setting m_+ = m_− = m_B yields:

    μ_B = μ_ref,m,B + RT ln[(γ_+ γ_−)(m_B / m°)²] (for ν = 2)

  • The key feature: the term (m_B / m°)² appears, not (m_B / m°) as for a nonelectrolyte.

  • This squared dependence arises because both ions contribute to the solute chemical potential, and each ion's molality equals m_B.

⚠️ Don't confuse with nonelectrolytes

  • For a nonelectrolyte solute, the chemical potential depends on the first power of molality.
  • For a symmetrical electrolyte, the second power reflects the dissociation into two ions.
  • Example: doubling the molality of a nonelectrolyte changes μ by RT ln(2); for a symmetrical electrolyte, the change is RT ln(2²) = 2RT ln(2).

📏 Mean ionic activity coefficient

📏 Definition and measurability

Mean ionic activity coefficient γ_± for a symmetrical electrolyte: γ_± = √(γ_+ · γ_−) (for ν = 2)

  • Individual ion activity coefficients γ_+ and γ_− cannot be measured experimentally.

  • However, their product γ_+ · γ_− can be evaluated, so γ_± is a measurable quantity.

  • Using γ_±, the chemical potential becomes:

    μ_B = μ_ref,m,B + RT ln[(γ_±)²(m_B / m°)²] (for ν = 2)

🔬 How to determine γ_±

The excerpt mentions three methods:

  1. Vapor pressure measurements: if the electrolyte is sufficiently volatile (e.g., HCl), measure the partial pressure of the equilibrated gas phase.
  2. Osmotic coefficients: Section 10.6 describes a general method.
  3. Galvanic cells: Section 14.5 explains how to evaluate γ_± from equilibrium cell potentials in favorable cases.

🧮 Activity of the solute

  • The activity a_m,B is defined by: μ_B = μ°_m,B + RT ln(a_m,B)

  • Comparing with the expression using γ_± gives:

    a_m,B = Γ_m,B (γ_±)²(m_B / m°)² (for ν = 2)

  • Γ_m,B is the pressure factor (defined by the standard-state pressure).

  • The activity depends on the square of both the mean ionic activity coefficient and the molality ratio.

🚫 Unmeasurable vs measurable quantities

🚫 What cannot be determined

The excerpt emphasizes several times that certain single-ion quantities cannot be measured:

  • Individual ion chemical potentials μ_+ and μ_− (absolute values)
  • Individual ion activities a_+ and a_−
  • Individual ion activity coefficients γ_+ and γ_−
  • Individual ion pressure factors Γ_+ and Γ_−

Why: there is no experimental way to measure the energy brought into the system by a single ion species in isolation, because ions always come with counterions to maintain electroneutrality.

✅ What can be determined

QuantityWhy it is measurable
μ_B − μ_ref,m,BDifference in solute chemical potential from a reference state
μ_B − μ°_m,BDifference from the standard state
γ_+ · γ_− (or γ_±)Product of ion activity coefficients, determined from solute properties
a_m,BSolute activity, from measurable chemical potential differences
  • The key insight: treating the electrolyte as a whole allows thermodynamic analysis, even though individual ion properties remain inaccessible.

🔍 Electric potential independence

  • The individual ion chemical potentials μ_+ and μ_− depend on the electric potential Φ of the solution phase.
  • However, for a symmetrical electrolyte, the sum z_+ + z_− = 0 (charges cancel).
  • Therefore, μ_B = μ_+(0) + μ_−(0) does not depend on Φ.
  • Don't confuse: the solute chemical potential is independent of Φ, but the ion chemical potentials are not.
62

Electrolytes in General

10.3 Electrolytes in General

🧭 Overview

🧠 One-sentence thesis

For non-symmetrical electrolytes (with more than two ions per formula unit), the solute activity and mean ionic activity coefficient follow the same principles as symmetrical electrolytes but with weighted geometric averages that account for the different numbers of cations and anions.

📌 Key points (3–5)

  • Extension to non-symmetrical electrolytes: formulas become more complicated than the symmetrical case but use the same reasoning, with stoichiometric coefficients ν₊ (cations) and ν₋ (anions) per formula unit.
  • Mean ionic activity coefficient definition: γ± is a geometric average of single-ion activity coefficients weighted by the number of each ion type (ν₊ and ν₋).
  • Activity depends on total ion count: solute activity is proportional to molality raised to the power ν (total ions), not just squared as in symmetrical cases.
  • Multisolute solutions: the relation μ_B = ν₊μ₊ + ν₋μ₋ holds for each solute even when multiple electrolytes share a common ion.
  • Common confusion: in multisolute solutions, ion molalities are not necessarily in the same stoichiometric ratio as in the pure solute substance because ions may come from multiple sources.

🔢 Stoichiometric coefficients and ion molalities

🔢 Defining the ion counts

The excerpt defines three key symbols for any electrolyte:

  • ν₊: number of cations per solute formula unit
  • ν₋: number of anions per solute formula unit
  • ν: the sum ν₊ + ν₋ (total ions per formula unit)

Example: For Al₂(SO₄)₃, ν₊ = 2, ν₋ = 3, and ν = 5.

💧 Relating ion molalities to solute molality

For a single electrolyte solution:

m₊ = ν₊ m_B
m₋ = ν₋ m_B

where m_B is the overall solute molality and m₊, m₋ are the cation and anion molalities.

  • This generalizes the symmetrical case where ν₊ = ν₋ = 1.
  • These relations assume complete dissociation.

⚗️ Chemical potential relationships

⚗️ Solute chemical potential from ion potentials

The chemical potential of the electrolyte solute is related to single-ion chemical potentials by:

μ_B = ν₊μ₊ + ν₋μ₋

  • This comes from the additivity rule for Gibbs energy: G = n_A μ_A + n_B μ_B = n_A μ_A + ν₊n_B μ₊ + ν₋n_B μ₋.
  • The excerpt shows that electrical neutrality (ν₊z₊ + ν₋z₋ = 0) ensures μ_B does not depend on the electric potential φ, even though individual ion potentials do.

📐 General expression for solute chemical potential

By combining the ion-molality relations with reference states, the excerpt derives:

μ_B = μ_B^ref + RT ln[(γ₊^ν₊ γ₋^ν₋)(m_B/m°)^ν]

where:

  • μ_B^ref is the chemical potential in the reference state (standard molality, infinite-dilution behavior, φ = 0)
  • γ₊ and γ₋ are single-ion activity coefficients
  • The term (m_B/m°)^ν shows the activity depends on the total ion count ν, not just the solute molality itself

🧮 Mean ionic activity coefficient

🧮 Definition for general electrolytes

The mean ionic activity coefficient is defined as a weighted geometric average:

γ± = (γ₊^ν₊ γ₋^ν₋)^(1/ν)

Equivalently:

γ±^ν = γ₊^ν₊ γ₋^ν₋

  • This generalizes the symmetrical case where ν = 2 and γ± = √(γ₊γ₋).
  • The weighting by ν₊ and ν₋ reflects the different numbers of cations and anions.

📊 Why γ± is measurable

  • We cannot measure γ₊ or γ₋ individually (no way to isolate single-ion contributions).
  • However, γ± is measurable because μ_B − μ_B^ref is measurable from experiments.
  • The excerpt notes: "We have no way of evaluating γ₊ or γ₋ individually, even if we know the value of γ±. For instance, we cannot assume that γ₊ and γ₋ are equal."

🔬 Solute activity expression

The activity of the electrolyte solute on a molality basis is:

a_{m,B} = (ν₊^ν₊ ν₋^ν₋) λ_{m,B} γ±^ν (m_B/m°)^ν

where λ_{m,B} is a pressure factor (close to 1 at low pressures).

Key features:

  • Activity is proportional to (m_B)^ν in the infinite-dilution limit, not (m_B)^1 as for nonelectrolytes.
  • The factor (ν₊^ν₊ ν₋^ν₋) is a constant for a given electrolyte stoichiometry.

🧪 Multisolute solutions

🧪 Chemical potential relations with common ions

The relation μ_B = ν₊μ₊ + ν₋μ₋ holds for each individual solute in a multisolute solution, even when solutes share a common ion.

Example from the excerpt: A solution of BaI₂ and CsI in water.

  • BaI₂ dissociates into Ba²⁺ and I⁻
  • CsI dissociates into Cs⁺ and I⁻
  • I⁻ is common to both solutes

The Gibbs energy can be written as:

  • G = n_A μ_A + n_B μ_B + n_C μ_C (using solute quantities)
  • G = n_A μ_A + n_B μ(Ba²⁺) + 2n_B μ(I⁻) + n_C μ(Cs⁺) + n_C μ(I⁻) (using ion quantities)

Comparing these gives:

  • μ_B = μ(Ba²⁺) + 2μ(I⁻)
  • μ_C = μ(Cs⁺) + μ(I⁻)

Note that μ(I⁻), the chemical potential of the common ion, appears in both relations.

⚠️ Ion molalities vs. stoichiometric ratios

In multisolute solutions, ion molalities are not necessarily in the same stoichiometric ratio as in the pure solute substance.

Example: For BaI₂ activity in the BaI₂ + CsI solution:

  • The activity expression uses m(Ba²⁺) and m(I⁻)
  • But Ba²⁺ and I⁻ are not in a 1:2 ratio because I⁻ also comes from CsI
  • γ± still refers to the mean ionic activity coefficient of dissolved BaI₂

The general activity relation for any solute B in a multisolute solution:

a_{m,B} = λ_{m,B} γ± (m₊/m°)^ν₊ (m₋/m°)^ν₋

where m₊ and m₋ are the actual ion molalities (from all sources), not necessarily ν₊m_B and ν₋m_B.

🔓 Incomplete dissociation

🔓 When complete dissociation fails

Some solutions do not have complete dissociation:

  • Ion pairs: closely associated ions of opposite charge
  • Weak electrolytes: equilibrium between ions and electrically-neutral molecules

In these cases, the relations m₊ = ν₊m_B and m₋ = ν₋m_B are no longer valid.

🔓 Stoichiometric activity coefficient

When dissociation is incomplete:

  • The expression μ_B = μ_B^ref + RT ln[(γ±^ν)(m_B/m°)^ν] can still be used
  • However, γ± no longer represents the geometric average of actual dissociated-ion activity coefficients
  • Instead, γ± is called the stoichiometric activity coefficient of the electrolyte

Don't confuse:

  • True mean ionic activity coefficient: geometric average of γ₊ and γ₋ for fully dissociated ions
  • Stoichiometric activity coefficient: an effective coefficient used when dissociation is incomplete; it accounts for both incomplete dissociation and non-ideal interactions
63

The Debye–Hückel Theory

10.4 The Debye–Hückel Theory

🧭 Overview

🧠 One-sentence thesis

The Debye–Hückel theory provides theoretical expressions for activity coefficients in electrolyte solutions by considering electrostatic interactions between ions and the non-random ion atmosphere that forms around each ion.

📌 Key points

  • What the theory does: provides theoretical expressions for single-ion and mean ionic activity coefficients based solely on electrostatic interactions between ions.
  • Why activity coefficients change: each ion is surrounded by a non-random "ion atmosphere" with a surplus of oppositely charged ions, creating a net attractive interaction that lowers activity coefficients below 1.
  • Two forms of the equation: the full Debye–Hückel equation (with adjustable parameter a) fits experimental data over a wider range; the limiting law (no adjustable parameters) works only at very low molality.
  • Ionic strength matters: the theory uses ionic strength (calculated from all ions in solution) because each ion interacts with the atmosphere created by all other ions, not just its own counterions.
  • Common confusion: in multisolute solutions, ionic strength includes all ions present, not just those from a single solute; ions may not be in stoichiometric ratios.

⚡ Electrostatic interactions and ion atmospheres

⚡ Why electrostatic interactions dominate

  • The theory considers only electrostatic interactions between ions.
  • These interactions are much stronger than those between uncharged molecules.
  • They also die off more slowly with distance compared to interactions between neutral molecules.

🌐 The ion atmosphere concept

  • If ion positions were completely random, electrostatic effects would cancel: each cation–cation repulsion would balance a cation–anion attraction, giving zero net effect.
  • Positions are not random: each cation has a surplus of anions in its immediate environment, and each anion has a surplus of neighboring cations.
  • Result: each ion experiences a net attractive interaction with the surrounding ion atmosphere.
  • This attraction lowers the activity coefficient below 1 as molality increases beyond the ideal-dilute range.

Example: A cation moving through solution is surrounded by more anions than cations on average; this non-random distribution creates an attractive environment that reduces the cation's effective activity.

📐 The Debye–Hückel equations

📐 Single-ion activity coefficient

The single-ion activity coefficient γ_i of ion i is given by: ln γ_i = −A_DH · z_i² · √I_m / (1 + B_DH · a · √I_m)

Where:

  • z_i: charge number of ion i (+1, −2, etc.)
  • I_m: ionic strength on a molality basis
  • A_DH and B_DH: functions of solvent properties and temperature
  • a: adjustable parameter equal to the mean effective distance of closest approach of other ions to ion i

📐 Mean ionic activity coefficient

For an electrolyte solute: ln γ_± = −A_DH · |z_+ · z_−| · √I_m / (1 + B_DH · a · √I_m)

  • z_+ and z_−: charge numbers of the cation and anion of the solute
  • The right side is negative at finite molalities and zero at infinite dilution.
  • Prediction: γ_± is less than 1 at finite molalities and approaches 1 at infinite dilution.

🔬 The limiting law

At infinite dilution: ln γ_± = −A_DH · |z_+ · z_−| · √I_m

  • Known as the Debye–Hückel limiting law.
  • Contains no adjustable parameters.
  • Agrees with experiment only at quite low molality (see Fig. 10.3, dotted curve).
  • The full equation with parameter a (dashed curve) agrees closely with experiment over a wider range.

Don't confuse: The limiting law is a simplified form valid only at very low concentration; the full equation includes the distance parameter a and works at higher concentrations.

🧪 Ionic strength and its calculation

🧪 Definition of ionic strength

Ionic strength I_m is defined by: I_m = (1/2) · Σ(all ions) m_j · z_j²

  • m_j: molality of ion j
  • z_j: charge number of ion j
  • The sum is over all ions in the solution, not just those from a single solute.

🧪 Why all ions matter

  • The departure of γ_+ and γ_− from the ideal-dilute value of 1 is caused by interaction of each ion with the ion atmosphere resulting from all other ions in the solution.
  • In multisolute solutions, you must include ions from every solute present.

🧪 Ionic strength for single electrolytes

For a binary solution of one completely dissociated electrolyte, the relationship between ionic strength and solute molality m_B depends on stoichiometry:

Electrolyte typeExampleRelationship
1:1NaCl, HClI_m = m_B
1:2 or 2:1Na₂SO₄, CaCl₂I_m = 3m_B
2:2MgSO₄I_m = 4m_B
1:3 or 3:1AlCl₃I_m = 6m_B
3:2 or 2:3Al₂(SO₄)₃I_m = 15m_B

The general relation: I_m = (1/2) · ν · |z_+ · z_−| · m_B, where ν is the total number of ions per formula unit.

📊 Agreement with experiment

📊 Performance at different concentrations

  • Figure 10.3 (aqueous HCl at 25°C) shows three curves:
    • Solid curve: experimental data
    • Dashed curve: full Debye–Hückel equation with a = 5 × 10⁻¹⁰ m—agrees closely with experiment at low molalities
    • Dotted curve: limiting law—agrees only at quite low molality

📊 Dependence on ionic strength

  • Figure 10.4 shows ln γ_± versus √I_m for HCl and CaCl₂.
  • Experimental curves have the limiting slopes predicted by the limiting law.
  • At low ionic strength, curves deviate significantly from the linear relations predicted by the limiting law.
  • The full equation fits experimental curves over a wider range of ionic strength.

📊 Solvent and temperature dependence

For water at 25°C:

  • A_DH = 1.1744 kg^(1/2) mol^(−1/2)
  • B_DH = 3.285 × 10⁹ m⁻¹ kg^(1/2) mol^(−1/2)

These constants are defined functions of solvent density, relative permittivity (dielectric constant), temperature, Avogadro constant, elementary charge, and electric constant.

🔄 Incomplete dissociation and stoichiometric coefficients

🔄 When dissociation is not complete

  • Some solutions contain ion pairs: closely associated ions of opposite charge.
  • "Weak" electrolytes establish equilibrium between ions and electrically-neutral molecules.
  • In these cases, the simple relations between solute molality and ion molalities (from earlier sections) are no longer valid.

🔄 Stoichiometric activity coefficient

  • When dissociation is incomplete, the expression for μ_B can still be used.
  • However, γ_± no longer has the physical significance of being the geometric average of the activity coefficients of the actual dissociated ions.
  • It is then called the stoichiometric activity coefficient of the electrolyte.

Don't confuse: The stoichiometric activity coefficient is a formal quantity used when dissociation is incomplete; it does not directly represent the activity coefficients of free ions.

🧬 Theoretical foundation

🧬 Derivation approach

  • Debye and Hückel derived the equation using a combination of:
    • Electrostatic theory
    • Statistical mechanical theory
    • Thermodynamics

🧬 Central ion and distribution

  • The derivation focuses on an individual ion (the "central ion") as it moves through solution.
  • Around this central ion, the time-average spatial distribution of any ion species j is not random, due to interaction with the central ion.
  • The distribution must be spherically symmetric about the central ion: a function only of distance r from the ion center.
  • Species i and j may be the same or different.
64

Derivation of the Debye–Hückel Equation

10.5 Derivation of the Debye–Hückel Equation

🧭 Overview

🧠 One-sentence thesis

The Debye–Hückel equation is derived by combining electrostatic theory, statistical mechanics, and thermodynamics to calculate how ion–ion interactions in solution affect the chemical potential of each ion species.

📌 Key points (3–5)

  • Core approach: the derivation focuses on a "central ion" and calculates the time-average distribution of surrounding ions (the "ion atmosphere") using the Boltzmann distribution.
  • Key assumption: the local concentration of ions around the central ion depends on electric potential and charge, and the potential is spherically symmetric about the central ion.
  • Calculating the interaction energy: the work to charge the central ion from zero to its actual charge, while surrounded by the ion atmosphere, equals the Gibbs energy change due to ion–ion interactions.
  • Common confusion: the electric potential at a point is split into two parts—one from the central ion alone (as if at infinite dilution) and one from all other ions (the ion atmosphere); don't confuse the total potential with just the central ion's contribution.
  • Final step: equating the reversible work to the chemical potential difference yields the Debye–Hückel equation for the activity coefficient.

🔬 Starting point and setup

🎯 The central ion and ion atmosphere

  • The derivation begins by selecting one ion of species i as the "central ion" and examining how it moves through the solution.
  • Around this central ion, the time-average spatial distribution of any ion species j (which may be the same as or different from i) is not random because of electrostatic interactions.
  • The distribution must be spherically symmetric about the central ion—it depends only on the distance r from the center.

🌐 What must be found

  • Two quantities must be determined as functions of r and must be mutually consistent:
    • The local concentration c'_j of ions of species j at distance r.
    • The electric potential ψ at distance r.
  • Both depend on ion charge and the distribution of all ions in solution.

⚡ The Boltzmann distribution and electric potential

📊 Boltzmann distribution for local concentration

The local concentration c'_j is given by the Boltzmann distribution: c'_j = c_j exp(−z_j eψ / kT), where z_j e is the charge of species j, ψ is the electric potential, k is the Boltzmann constant (k = R / N_A), and c_j is the macroscopic concentration.

  • As r becomes large, ψ approaches zero and c'_j approaches the bulk concentration c_j.
  • As temperature T increases, c'_j at a fixed r approaches c_j because thermal energy randomizes the distribution.
  • Approximation used: Debye and Hückel expanded the exponential in powers of 1/T and kept only the first two terms: c'_j ≈ c_j (1 − z_j eψ / kT).

🔋 Electric potential function

The electric potential consistent with the ion distribution and overall electroneutrality is:

ψ = (z_i e / 4πε_r ε_0 r) exp[−κ(a − r)] / (1 + κa)

  • Here κ is defined by κ² = 2N²_A e² I_c / (ε_r ε_0 RT), where I_c is the ionic strength on a concentration basis: I_c = (1/2) Σ_i c_i z²_i.
  • The potential ψ at any point is assumed to be the sum of two contributions:
    1. The potential the central ion would cause at infinite dilution: z_i e / (4πε_r ε_0 r).
    2. The potential due to all other ions (the ion atmosphere): ψ'.

🧩 Splitting the potential

  • The potential from the ion atmosphere is:

ψ' = (z_i e / 4πε_r ε_0 r) [exp(−κ(a − r)) / (1 + κa) − 1]

  • This expression is valid for distances r ≥ a, where a is the distance of closest approach of other ions to the central ion.
  • For r < a, ψ' is constant and equal to ψ'(a) = (z_i e / 4πε_r ε_0) κ / (1 + κa).
  • Don't confuse: ψ is the total potential; ψ' is only the contribution from the ion atmosphere, not from the central ion itself.

🔧 Calculating the interaction energy

⚙️ Interaction energy of the central ion with its atmosphere

  • The interaction energy between the central ion and the surrounding ion atmosphere is the product of the central ion's charge z_i e and the potential ψ'(a) at the distance of closest approach.
  • This gives the energy: z_i e · ψ'(a) = (z²_i e² / 4πε_r ε_0) κ / (1 + κa).

🔄 Reversible charging process

  • The final step calculates the work of a hypothetical reversible process:
    • The surrounding ions remain in their final distribution.
    • The charge of the central ion gradually increases from zero to its actual value z_i e.
  • Let αz_i e be the charge at each stage, where α is a fractional advancement changing from 0 to 1.
  • The work w' due to the interaction with the ion atmosphere is obtained by integrating ψ'(a) over the charge:

w' = ∫[α=0 to α=1] [(αz_i e / 4πε_r ε_0) κ / (1 + κa)] d(αz_i) = (z²_i e² / 8πε_r ε_0) κ / (1 + κa)

🧪 Connecting work to Gibbs energy

  • In a reversible process at constant T and p, the infinitesimal Gibbs energy change is dG = −S dT + V dp + δw'.
  • Therefore, this reversible nonexpansion work equals the Gibbs energy change.
  • The Gibbs energy change per amount of species i is:

w' N_A = (z²_i e² N_A / 8πε_r ε_0) κ / (1 + κa)

  • This quantity is ΔG / n_i for the process in which a solution of fixed composition changes from a hypothetical state lacking ion–ion interactions to the real state with ion–ion interactions present.

🎯 Deriving the activity coefficient

🔗 Relating Gibbs energy to chemical potential

  • ΔG / n_i can be equated to the difference in chemical potentials of species i in the final and initial states.
  • If the chemical potential without ion–ion interactions is taken to be that for ideal-dilute behavior on a molality basis:

μ_i = μ°_ref,m,i + RT ln(m_i / m°)

  • Then:

−(z²_i e² N_A / 8πε_r ε_0) κ / (1 + κa) = μ_i − [μ°_ref,m,i + RT ln(m_i / m°)] = RT ln γ_m,i

🧮 Final substitution

  • In a dilute solution, the concentration c_i can be set equal to ρ_A m_i (where ρ_A is the solvent density), and the ionic strength I_c can be set equal to ρ_A I_m.
  • Substituting these approximations into the expression for κ and solving for ln γ_m,i yields Equation 10.4.1, the Debye–Hückel equation.

📌 Summary of the derivation logic

StepWhat is doneWhy
1. Define central ionFocus on one ion and its surroundingsSimplifies the problem to spherical symmetry
2. Boltzmann distributionRelate local ion concentration to potentialStatistical mechanics gives the distribution
3. Electric potentialFind ψ consistent with the distributionElectrostatics and electroneutrality constrain ψ
4. Reversible chargingCalculate work to charge the central ionWork equals Gibbs energy change
5. Equate to chemical potentialRelate Gibbs energy to activity coefficientThermodynamics connects energy to activity
65

Mean Ionic Activity Coefficients from Osmotic Coefficients

10.6 Mean Ionic Activity Coefficients from Osmotic Coefficients

🧭 Overview

🧠 One-sentence thesis

The mean ionic activity coefficient of an electrolyte can be calculated from osmotic coefficient data using an integration procedure that avoids numerical difficulties at infinite dilution by starting from a low molality where the activity coefficient is known or can be estimated.

📌 Key points

  • What is being calculated: the mean ionic activity coefficient (γ±) of a strong electrolyte or stoichiometric activity coefficient of a partially dissociated electrolyte from osmotic coefficient (φₘ) measurements.
  • Data sources: osmotic coefficients are commonly obtained from isopiestic measurements or freezing-point depression experiments.
  • Key challenge: direct integration from zero molality is difficult because the integrand approaches negative one as molality approaches zero (a problem unique to electrolytes, not nonelectrolytes).
  • Solution strategy: split the integration into two parts—use the Debye–Hückel equation or measurements to estimate γ± at a low molality (m″ᵦ), then integrate numerically from that point to the desired molality.
  • Common confusion: the osmotic coefficient definition for electrolytes differs from nonelectrolytes—the sum of individual ion molalities is replaced by νmᵦ (total ion molality assuming complete dissociation).

🔬 Osmotic coefficient definition for electrolytes

🔬 How φₘ is defined for electrolyte solutions

For a binary electrolyte solution: φₘ = −μₐ/(RTMₐνmᵦ)

  • The key difference from nonelectrolyte solutions: the sum of individual ion molalities (Σmᵢ) is replaced by νmᵦ, where ν is the stoichiometric coefficient and mᵦ is the molality assuming complete dissociation.
  • This definition applies to both strong electrolytes and those that do not dissociate completely.
  • The osmotic coefficient relates the chemical potential of the solvent (μₐ) to the total ion molality.

📊 Where φₘ data comes from

Two common experimental methods provide osmotic coefficient values:

MethodReference in textWhat it measures
Isopiestic methodSec. 9.6.4Vapor pressure equilibrium between solutions
Freezing-point depressionSec. 12.2Colligative property related to solvent activity

🧮 Derivation of the integration formula

🧮 Starting from thermodynamic relationships

The derivation parallels the approach for nonelectrolyte solutions but uses electrolyte-specific definitions:

  1. Solve for μₐ from the osmotic coefficient definition and take its differential at constant T and p:

    • dμₐ = RTMₐ(φₘ dmᵦ + mᵦ dφₘ)
  2. Express dμᵦ from the mean ionic activity coefficient (Eq. 10.3.9):

    • dμᵦ = RT ν(d ln γ± + dmᵦ/mᵦ)
  3. Apply the Gibbs–Duhem equation (nₐ dμₐ + nᵦ dμᵦ = 0) with the substitution nₐMₐ = nᵦ/mᵦ.

📐 The resulting differential equation

After substitution and rearrangement:

  • d ln γ± = dφₘ + (φₘ − 1)/mᵦ dmᵦ

This differential equation connects changes in the mean ionic activity coefficient to changes in the osmotic coefficient.

🎯 Integration to the final formula

Integrating from mᵦ = 0 to any desired molality m′ᵦ gives:

  • ln γ±(m′ᵦ) = φₘ(m′ᵦ) − 1 + ∫₀^(m′ᵦ) [(φₘ − 1)/mᵦ] dmᵦ

Important observation: The right side has the same form as the expression for ln γₘ;ᵦ for a nonelectrolyte, but the integrand behaves differently.

⚠️ The numerical integration problem

⚠️ Why direct integration fails

  • The difficulty: The integrand (φₘ − 1)/mᵦ approaches −1 as mᵦ approaches zero.
  • This behavior makes numerical integration starting at mᵦ = 0 unreliable or impossible.
  • Key distinction: This problem does not exist for nonelectrolyte solutions—it is specific to electrolytes due to long-range ionic interactions.

Example: If you try to evaluate the integral numerically from zero, the function becomes singular (undefined or unstable) at the lower limit.

⚠️ Why electrolytes behave differently

The excerpt notes that "this difficulty does not exist when the solute is a nonelectrolyte," implying that the limiting behavior of φₘ as concentration approaches zero differs fundamentally between electrolytes and nonelectrolytes due to the nature of ionic interactions.

🔧 The two-part integration solution

🔧 Splitting the integral

To avoid the numerical problem, split the integration range:

  • ∫₀^(m′ᵦ) [(φₘ − 1)/mᵦ] dmᵦ = ∫₀^(m″ᵦ) [(φₘ − 1)/mᵦ] dmᵦ + ∫_(m″ᵦ)^(m′ᵦ) [(φₘ − 1)/mᵦ] dmᵦ

Where m″ᵦ is a carefully chosen low molality that satisfies two conditions:

  1. Osmotic coefficient data (φₘ) is available at this molality
  2. The mean ionic activity coefficient γ± can either be measured or estimated (e.g., from the Debye–Hückel equation)

🔧 Eliminating the problematic integral

Write the integration formula at the low molality m″ᵦ:

  • ln γ±(m″ᵦ) = φₘ(m″ᵦ) − 1 + ∫₀^(m″ᵦ) [(φₘ − 1)/mᵦ] dmᵦ

Rearrange to isolate the problematic integral:

  • ∫₀^(m″ᵦ) [(φₘ − 1)/mᵦ] dmᵦ = ln γ±(m″ᵦ) − φₘ(m″ᵦ) + 1

Substitute this back into the split integral to eliminate the term that cannot be evaluated numerically.

🎯 Final working equation

The practical formula for calculating γ± at any molality m′ᵦ:

  • ln γ±(m′ᵦ) = φₘ(m′ᵦ) − φₘ(m″ᵦ) + ln γ±(m″ᵦ) + ∫_(m″ᵦ)^(m′ᵦ) [(φₘ − 1)/mᵦ] dmᵦ

Key advantage: The integral now runs from m″ᵦ (a small but nonzero value) to m′ᵦ, avoiding the singularity at zero and making numerical integration straightforward.

📋 Practical application procedure

📋 Step-by-step calculation

To evaluate γ± at a desired molality:

  1. Choose a low reference molality m″ᵦ where:

    • Osmotic coefficient φₘ(m″ᵦ) is available from experimental data
    • Activity coefficient γ±(m″ᵦ) can be estimated (often using the Debye–Hückel equation)
  2. Collect osmotic coefficient data φₘ(mᵦ) from m″ᵦ to the target molality m′ᵦ.

  3. Evaluate the integral numerically: ∫_(m″ᵦ)^(m′ᵦ) [(φₘ − 1)/mᵦ] dmᵦ using the collected data.

  4. Apply the final equation to calculate ln γ±(m′ᵦ).

📋 Example from the problems

Problem 10.2 illustrates the procedure for Na₂SO₄:

  • Use Debye–Hückel equation (Eq. 10.4.7) with parameter a = 3.0 × 10⁻¹⁰ m to estimate γ± at low molality
  • Use osmotic coefficient data from freezing-point measurements (converted to 298.15 K)
  • Calculate γ± at mᵦ = 0.15 mol/kg using the integration formula

Don't confuse: The parameter "a" in the Debye–Hückel equation is an ion-size parameter (in meters), not the activity or activity coefficient.

66

Mixing Processes

11.1 Mixing Processes

🧭 Overview

🧠 One-sentence thesis

Mixing processes are characterized by changes in Gibbs energy and other thermodynamic properties, with ideal mixtures showing predictable behavior (negative Gibbs energy change, positive entropy change, zero enthalpy and volume changes) while real mixtures exhibit deviations quantified by excess properties.

📌 Key points (3–5)

  • Gibbs energy of mixing: Always negative for ideal mixtures at any composition, driving spontaneous mixing.
  • Ideal vs. real mixtures: Ideal mixtures have zero enthalpy and volume changes upon mixing; real mixtures show deviations captured by excess quantities.
  • Entropy increase in ideal mixing: Results from volume expansion of each component, not from intermingling of different substances per se.
  • Common confusion: The entropy increase in ideal gas mixing comes from each gas occupying a larger volume, not fundamentally from mixing different substances—isothermal expansion of unmixed gases to the same volumes gives identical entropy change.
  • Phase separation: Occurs when the Gibbs energy of mixing curve becomes concave downward, allowing two liquid phases to coexist at lower total Gibbs energy.

🧪 Thermodynamics of mixing

🧪 Gibbs energy change

Gibbs energy of mixing: The difference between the Gibbs energy of the mixture and the sum of Gibbs energies of the pure components at the same temperature and pressure.

The general expression is:

  • ΔG(mix) = sum over all components i of: n_i × (μ_i - μ°_i)
  • Where μ_i is the chemical potential in the mixture and μ°_i is for the pure substance
  • The molar quantity divides by total amount: ΔG_m(mix) = sum of x_i × (μ_i - μ°_i)

For ideal mixtures specifically:

  • ΔG^id_m(mix) = RT × sum of x_i × ln(x_i)
  • Since each mole fraction is less than 1, each logarithm is negative
  • Therefore ΔG^id_m(mix) is always negative for any composition
  • This negative value drives spontaneous mixing

🔥 Enthalpy and internal energy

Ideal mixtures:

  • ΔH^id_m(mix) = 0
  • ΔU^id_m(mix) = 0
  • Mixing is an "athermal" process—no heat transfer needed to maintain constant temperature

Why this matters:

  • Example: If you mix two ideal liquids at constant temperature and pressure, the container neither heats up nor cools down
  • The process requires no energy input or removal beyond maintaining pressure

📦 Volume and entropy changes

Volume:

  • ΔV^id_m(mix) = 0 for ideal mixtures
  • The mixture volume equals the sum of pure component volumes at the same T and p
  • Don't confuse: Water and methanol show volume change upon mixing, proving they form a nonideal mixture despite mixing in all proportions

Entropy:

  • ΔS^id_m(mix) = -R × sum of x_i × ln(x_i)
  • This quantity is always positive (since ln(x_i) is negative)
  • Note: Some nonideal mixtures can have negative entropy of mixing (example given: diethylamine and water at 322 K has ΔS_m = -8.8 J K⁻¹ mol⁻¹)

🎯 Understanding ideal gas mixing

🎯 The volume principle

The excerpt emphasizes a key insight about ideal gas mixing:

The entropy of an ideal gas mixture equals the sum of the entropies of the unmixed pure ideal gases, each pure gas having the same temperature and occupying the same volume as in the mixture.

What this means:

  • When ideal gases mix at constant T and p, each gas expands to fill the total volume
  • The entropy increase comes from this volume expansion, not from intermingling different substances
  • Example: Gas A initially in volume V₁(A) and gas B in volume V₁(B), both at pressure p, mix to form mixture of volume V₂ = V₁(A) + V₁(B)
  • Each gas now occupies volume V₂ instead of its initial smaller volume

🔄 Reversible mixing with semipermeable pistons

The excerpt describes a thought experiment:

  • Use two pistons: one permeable to A but not B, the other permeable to B but not A
  • Move the pistons apart reversibly and isothermally
  • Gas A exerts no net force on its piston (transfer equilibrium maintained)
  • Gas B exerts no net force on its piston
  • Total reversible work: w = n_A RT ln(y_A) + n_B RT ln(y_B)
  • Entropy change: ΔS = -w/T (since q = -w for isothermal process with ΔU = 0)

Key insight:

  • Isothermal expansion of both pure gases to volume V₂ without mixing gives the same entropy change
  • This proves the entropy increase is fundamentally about volume expansion, not about mixing different substances

⚠️ Common misconception addressed

Don't confuse: Removing a partition between two subsystems of the same pure ideal gas produces zero entropy change, even though molecules intermingle. The reason is that there is no macroscopic change of state—the same substance at the same T and p throughout.

🧬 Molecular models and excess properties

🧬 Quasicrystalline lattice model

The excerpt presents a molecular model for liquid mixtures:

Assumptions:

  • Each molecule is surrounded by nearest neighbors
  • Number of A and B neighbors around any molecule is proportional to mole fractions x_A and x_B
  • Molecules have similar sizes and shapes
  • Random mixing in the mixture

Interaction energies:

  • k_AA: interaction energy per mole for A-A pairs
  • k_BB: interaction energy per mole for B-B pairs
  • k_AB: interaction energy per mole for A-B pairs

Condition for ideal mixture:

  • Requires k_AB = (k_AA + k_BB)/2
  • In words: An A-B interaction must equal the average of an A-A interaction and a B-B interaction
  • When this condition holds, ΔU(mix) = 0

When deviations occur:

  • ΔU(mix) = (n_A × n_B / n) × (2k_AB - k_AA - k_BB)
  • If k_AB is less negative than the average, mixing is endothermic
  • If k_AB is more negative than the average, mixing is exothermic

📊 Excess quantities

Excess quantity X^E: The difference between the actual property value of a real mixture and the value for a hypothetical ideal mixture at the same temperature, pressure, and composition.

Key relationships:

  • X^E_m = ΔX_m(mix) - ΔX^id_m(mix)
  • For specific properties:
    • H^E_m = ΔH_m(mix) (since ΔH^id_m(mix) = 0)
    • U^E_m = ΔU_m(mix)
    • V^E_m = ΔV_m(mix)
    • G^E_m = ΔG_m(mix) - RT × sum of x_i ln(x_i)
    • S^E_m = ΔS_m(mix) + R × sum of x_i ln(x_i)

Connection to activity coefficients:

  • G^E_m = RT × sum of x_i ln(γ_i)
  • Where γ_i is the activity coefficient of component i
  • This provides a direct link between excess Gibbs energy and deviations from ideal behavior

📐 Redlich-Kister series

For real binary mixtures, a flexible empirical expression:

  • G^E_m = x_A × x_B × [a + b(x_A - x_B) + c(x_A - x_B)² + ...]
  • Parameters a, b, c depend on T and p but not composition
  • Must satisfy: G^E_m = 0 when x_A = 0 or x_B = 0
  • Two-parameter version: G^E_m = x_A × x_B × [a + b(x_A - x_B)]

🌊 Phase separation in liquid mixtures

🌊 Condition for phase separation

A binary liquid mixture can spontaneously separate into two liquid layers when:

  • The curve of ΔG_m(mix) versus x_A has a portion that is concave downward (negative curvature)
  • This requires at least two inflection points on the curve

🔍 Equilibrium between two liquid phases

Graphical method:

  • Plot ΔG_m(mix) versus x_A
  • Draw a common tangent to the curve at two points (labeled α and β)
  • The tangent intercepts give (μ_A - μ°_A) and (μ_B - μ°_B)
  • Because the tangent is common, both phases have the same chemical potentials
  • Transfer equilibrium conditions satisfied: μ^α_A = μ^β_A and μ^α_B = μ^β_B

Thermodynamic driving force:

  • Any single-phase composition between x^α_A and x^β_A is unstable
  • Separating into two phases lowers the total Gibbs energy
  • The decrease in G/n is shown by the vertical distance from the curve to the common tangent

📈 Activity and miscibility gaps

When phase separation occurs:

  • The activity a_A shows a maximum in its plot versus x_A
  • The portion of the curve between coexisting phase compositions represents unstable states
  • Both phases have the same activity a_A (indicated by horizontal line on activity plot)
  • The difference in compositions is called a "miscibility gap"

Example using Redlich-Kister model:

  • Curve 1 (a = b = 0): Ideal mixture, no phase separation
  • Curve 2 (a/RT = 1.8, b/RT = 0.36): Positive G^E_m, no phase separation, monotonic activity increase
  • Curve 3 (a/RT = 2.4, b/RT = 0.48): Larger positive G^E_m, phase separation occurs, activity maximum appears

Don't confuse: The compositions of coexisting phases are not at the inflection points, nor necessarily at the local minima of the ΔG_m(mix) curve—they are at the points where a common tangent touches the curve.

67

The Advancement and Molar Reaction Quantities

11.2 The Advancement and Molar Reaction Quantities

🧭 Overview

🧠 One-sentence thesis

The advancement variable tracks the progress of a chemical reaction in a closed system and allows us to express how extensive properties change with reaction progress through molar reaction quantities.

📌 Key points (3–5)

  • What advancement (ξ) measures: the amount (in moles) by which a reaction has progressed forward from initial conditions in a closed system.
  • How amounts change: in a closed system, the amount of each species changes according to its stoichiometric number times the advancement: n_i = n_i,0 + ν_i ξ.
  • Molar reaction quantity Δ_r X: the rate at which an extensive property X changes with advancement at constant T and p; defined as the sum of stoichiometric numbers times partial molar quantities.
  • Common confusion: Δ_r X (differential, a property of a state) vs ΔX_m(rxn) (integral, ratio of finite changes)—they are equal only when partial molar quantities remain constant during the reaction.
  • Standard molar reaction quantities: when all species remain in standard states (unit activity), Δ_r X° depends only on temperature, not on pressure or advancement.

🔢 The advancement variable

🔢 Definition and meaning

Advancement (extent of reaction), ξ: the amount by which the reaction defined by the reaction equation has advanced in the forward direction from specified initial conditions.

  • Dimensions: amount of substance (unit: mole).
  • Tracks reaction progress in a closed system where amounts change only due to the reaction itself.
  • Example: for N₂ + 3 H₂ → 2 NH₃, if ξ = 1 mol, then 1 mole of N₂ has reacted, 3 moles of H₂ have been consumed, and 2 moles of NH₃ have been produced.

📐 How amounts depend on advancement

For the ammonia synthesis reaction N₂ + 3 H₂ → 2 NH₃, the amounts at any stage are:

  • n_N₂ = n_N₂,0 − ξ
  • n_H₂ = n_H₂,0 − 3ξ
  • n_NH₃ = n_NH₃,0 + 2ξ

Why these relations work:

  • The stoichiometric coefficients tell us the ratio in which species are consumed or produced.
  • When ξ increases by 1 mol, H₂ decreases by 3 mol (hence the −3ξ term).
  • Reactants have negative signs (amounts decrease); products have positive signs (amounts increase).

🔄 Infinitesimal changes

Taking differentials of the amount relations:

  • d n_N₂ = −d ξ
  • d n_H₂ = −3 d ξ
  • d n_NH₃ = +2 d ξ

Key insight: In a closed system, changes in amounts are not independent—they are all linked through d ξ.

🧮 Stoichiometric numbers and general formulation

🧮 Stoichiometric numbers ν_i

Stoichiometric number ν_i: a dimensionless quantity taken as negative for a reactant and positive for a product.

  • General stoichiometric relation: 0 = Σ_i ν_i A_i
  • For N₂ + 3 H₂ → 2 NH₃, we write: 0 = −N₂ − 3 H₂ + 2 NH₃
    • ν_N₂ = −1
    • ν_H₂ = −3
    • ν_NH₃ = +2

Don't confuse: Stoichiometric number vs stoichiometric coefficient. The number includes the sign (negative for reactants); the coefficient is just the magnitude.

📊 General amount relation

For any species i in a closed system:

  • n_i = n_i,0 + ν_i ξ
  • d n_i = ν_i d ξ

This holds for any chemical process, not just ammonia synthesis.

🔬 Molar reaction quantities

🔬 Definition of molar reaction quantity Δ_r X

Molar reaction quantity Δ_r X: the rate at which extensive property X changes with advancement at constant T and p.

Two equivalent definitions:

  1. As a partial derivative: Δ_r X = (∂X/∂ξ)_T,p
  2. As a sum over species: Δ_r X = Σ_i ν_i X_i

where X_i is the partial molar quantity of species i.

Example for enthalpy in ammonia synthesis:

  • Δ_r H = −H_N₂ − 3H_H₂ + 2H_NH₃
  • This is the molar reaction enthalpy (or molar enthalpy of reaction).

🧪 Physical interpretation

The excerpt provides a helpful thought experiment for understanding Δ_r H:

Imagined process at constant T and p in an open system:

  • Remove an infinitesimal amount d n of N₂
  • Remove 3 times this amount of H₂
  • Add 2 times this amount of NH₃
  • Total enthalpy change: d H = (−H_N₂ − 3H_H₂ + 2H_NH₃) d n

Why this matches the closed system:

  • The net change in state is equivalent to an advancement d ξ = d n in a closed system.
  • Therefore d H/d ξ in the closed system equals the combination of partial molar enthalpies.

📝 Dependence on how the equation is written

Important: The value of Δ_r X depends on how you write the reaction equation.

Example: If you change N₂ + 3 H₂ → 2 NH₃ to (1/2) N₂ + (3/2) H₂ → NH₃, then Δ_r H is halved.

This is because the stoichiometric numbers are halved, so the sum Σ_i ν_i X_i is halved.

🔀 Total differential with advancement

For a closed system with T, p, and ξ as independent variables:

d X = (∂X/∂T)_p,ξ d T + (∂X/∂p)_T,ξ d p + Δ_r X d ξ

  • The first two terms: changes due to temperature and pressure.
  • The third term: change due to reaction progress.
  • The subscript ξ on partial derivatives indicates amounts of all species are held constant (equivalent to the earlier notation {n_i}).

⚖️ Integral vs differential reaction quantities

⚖️ Two kinds of molar reaction quantities

TypeSymbolDefinitionMeaning
Molar integralΔX_m(rxn)ΔX(rxn)/ΔξRatio of finite differences between final and initial states at same T and p
Molar differentialΔ_r X(∂X/∂ξ)_T,p = Σ_i ν_i X_iRate of change of X with ξ at constant T and p; a property of the system in a given state

Common confusion: Both use the Δ symbol, but they are different concepts.

  • ΔX_m(rxn) is calculated from two states.
  • Δ_r X is a property at a single state (though it may vary as ξ changes).

🔁 When are they equal?

If partial molar quantities X_i remain constant for each species as the process advances at constant T and p, then:

  • Δ_r X is also constant.
  • X is a linear function of ξ.
  • ΔX_m(rxn) = Δ_r X for any finite change.

Example where they are equal:

  • Reactions in ideal gas mixtures, ideal condensed-phase mixtures, or ideal-dilute solutions.
  • In these systems, partial molar enthalpy H_i is independent of composition at constant T and p.
  • Therefore Δ_r H is constant, H is linear in ξ, and Δ_r H = ΔH_m(rxn).

Example where they differ:

  • Entropy in the same ideal gas mixture.
  • Partial molar entropy S_i depends on composition (even in ideal mixtures).
  • S is a nonlinear function of ξ.
  • The slope (Δ_r S) changes as the reaction advances.
  • ΔS_m(rxn) approaches Δ_r S only in the limit as Δξ → 0.

🌟 Standard molar reaction quantities

🌟 Definition and notation

Standard molar reaction quantity Δ_r X°: the molar reaction quantity when each reactant and product remains in its standard state of unit activity at constant temperature.

Formula: Δ_r X° = Σ_i ν_i X°_i

where X°_i is the standard molar quantity of species i.

🎯 Key properties

1. Depends only on temperature:

  • Standard-state conditions imply each species is in a separate phase of constant defined composition and constant pressure p°.
  • Therefore Δ_r X° is a function only of T, not of p or ξ.

2. Integral and differential are identical: ΔX°_m(rxn) = Δ_r X°

Why: Since standard conditions fix the state of each species, partial molar quantities are constant, so the integral and differential quantities must be equal.

📌 Notation summary

The excerpt emphasizes the operator interpretation: "Δ_r" can be thought of as an operator that acts on a property.

  • Δ_r H: molar differential reaction enthalpy
  • Δ_vap H°: standard molar enthalpy of vaporization
  • Δ_r G°: standard molar Gibbs energy of a reaction
  • Subscripts after Δ indicate the type of process: "r" (reaction in general), "vap" (vaporization), "sub" (sublimation), "fus" (fusion), "trs" (transition).

🔥 Connection to heat (brief preview)

🔥 Molar reaction enthalpy and heat transfer

The excerpt begins to connect molar reaction enthalpy to heat:

  • At constant pressure with expansion work only: d H = δq
  • For Δ_r H = (∂H/∂ξ)_T,p, the process is at both constant pressure and constant temperature.
  • (The excerpt cuts off here, but it sets up the relationship between Δ_r H and heat transfer in isothermal, isobaric reactions.)
68

11.3 Molar Reaction Enthalpy

11.3 Molar Reaction Enthalpy

🧭 Overview

🧠 One-sentence thesis

Molar reaction enthalpy quantifies the heat transferred during a chemical reaction at constant temperature and pressure, and its standard value can be calculated from tabulated formation enthalpies using Hess's law.

📌 Key points (3–5)

  • What molar reaction enthalpy measures: the heat transferred per unit advancement during a reaction at constant temperature and pressure (expansion work only).
  • Exothermic vs endothermic: negative molar reaction enthalpy means heat flows out (exothermic); positive means heat flows in (endothermic).
  • Hess's law principle: enthalpy is a state function, so the total enthalpy change is independent of path and equals the sum of enthalpies for any sequence of steps with the same net result.
  • Common confusion: molar reaction enthalpy equals heat transferred only when there is no nonexpansion work (e.g., electrical work); with nonexpansion work, enthalpy change and heat differ.
  • Temperature dependence: reaction enthalpy changes with temperature according to the Kirchhoff equation, which depends on the heat capacity difference between products and reactants.

🔥 Heat transfer and reaction types

🔥 Relationship between molar reaction enthalpy and heat

Molar reaction enthalpy (Δ_r H): the partial derivative of enthalpy with respect to advancement at constant temperature and pressure, equal to the sum over all species i of (stoichiometric number_i times partial molar enthalpy_i).

  • At constant temperature and pressure with expansion work only, the enthalpy change equals the heat transferred: dH = δq.
  • For molar reaction enthalpy: Δ_r H = δq / dξ (where ξ is advancement).
  • Why this matters: this equality holds only when there is no nonexpansion work (w' = 0).
  • Don't confuse: when nonexpansion work is present (e.g., in a galvanic cell), the heat transferred for a given advancement differs from the molar reaction enthalpy, even though Δ_r H remains the same.
  • Example: the same reaction in a galvanic cell vs. a reaction vessel has the same Δ_r H, but the heats may differ in magnitude or even sign.

❄️ Exothermic and endothermic reactions

Reaction typeSign of Δ_r HHeat flow (constant T, p)Temperature change (insulated, constant p)
ExothermicNegativeOut of systemIncreases
EndothermicPositiveInto systemDecreases
  • These definitions apply to all chemical processes at constant temperature and pressure, not just reactions.

📊 Standard molar enthalpies

📊 Standard molar reaction enthalpy

Standard molar reaction enthalpy (Δ_r H°): the molar reaction enthalpy when each reactant and product remains in its standard state of unit activity at constant temperature.

  • Standard molar reaction enthalpy is identical to the molar integral reaction enthalpy under standard state conditions.
  • Key property: depends only on temperature T, not on pressure p or advancement ξ, because standard-state conditions fix composition and pressure.
  • Practical assumption: unless experimental pressure is much greater than standard pressure p°, a reaction is exothermic if Δ_r H° is negative and endothermic if Δ_r H° is positive (because partial molar enthalpies depend only mildly on pressure).

🧱 Formation reactions and reference states

Formation reaction: the reaction in which a substance in a given physical state is formed from its constituent elements in their reference states at the same temperature.

Reference state of an element: usually the standard state of the element in the allotropic form and physical state that is stable at the given temperature and standard pressure.

Reference states at 298.15 K:

  • H₂, N₂, O₂, F₂, Cl₂, noble gases: ideal gas at 1 bar
  • Br₂ and Hg: liquid at 1 bar
  • P: crystalline white phosphorus at 1 bar (exception: red phosphorus is stable but not well characterized)
  • All other elements: stable crystalline allotrope at 1 bar
  • Example: carbon's reference state is graphite, not diamond, at 298.15 K and 1 bar.

🏗️ Standard molar enthalpy of formation

Standard molar enthalpy of formation (Δ_f H°): the enthalpy change per amount of substance produced in the formation reaction of the substance in its standard state.

  • Depends only on temperature T.
  • By definition: Δ_f H° for the reference state of an element is zero.
  • Example: for gaseous methyl bromide at 298.15 K, Δ_f H° is the molar reaction enthalpy of:
    • C(s, graphite, p°) + (3/2) H₂(ideal gas, p°) + (1/2) Br₂(l, p°) → CH₃Br(ideal gas, p°)

🧮 Hess's law and calculations

🧮 Hess's law principle

Hess's law: because enthalpy is a state function, ΔH for a given change of state is independent of path and equals the sum of ΔH values for any sequence of changes whose net result is the given change.

  • This principle applies to any state function, not just enthalpy.
  • How to use it: break a reaction into steps whose enthalpies are known, then sum them.

🔢 Calculating standard molar reaction enthalpy

Formula from Hess's law:

  • Δ_r H° = sum over all species i of (stoichiometric number_i times Δ_f H°(i))
  • In words: standard molar reaction enthalpy equals the sum of standard molar enthalpies of formation of products minus the sum for reactants, each multiplied by its stoichiometric coefficient.
  • Remember: stoichiometric numbers are negative for reactants and positive for products.

Example from the excerpt:

  • Combustion of graphite: C(s, graphite) + O₂(g) → CO₂(g), Δ_r H° = -393.51 kJ/mol
  • Combustion of CO: CO(g) + (1/2) O₂(g) → CO₂(g), Δ_r H° = -282.98 kJ/mol
  • Formation of CO: C(s, graphite) + (1/2) O₂(g) → CO(g)
  • By Hess's law: Δ_f H°(CO, g, 298.15 K) = (-393.51 + 282.98) kJ/mol = -110.53 kJ/mol
  • Why this works: the formation of CO is the first reaction minus the second reaction; direct measurement would be impractical because some CO₂ would form.

💧 Formation enthalpies for solutes and ions

For solutes in solution:

  • The formation reaction does not include forming the solvent from its elements.
  • Instead, the solute combines with the needed amount of pure liquid solvent.
  • For standard state: the amount of water needed is very large to achieve infinite-dilution behavior.
  • Example: formation of aqueous sucrose: 12 C(s, graphite) + 11 H₂(g) + (11/2) O₂(g) → C₁₂H₂₂O₁₁(aq)

For ions in solution:

  • No ordinary reaction produces a single ion without other species.
  • Solution: assign an arbitrary reference value to one ion, then build a consistent set.
  • Reference ion: aqueous hydrogen ion, Δ_f H°(H⁺, aq) = 0 at all temperatures.
  • Example: for the formation reaction (1/2) H₂(g) + (1/2) Cl₂(g) → H⁺(aq) + Cl⁻(aq), with measured Δ_r H° = -167.08 kJ/mol, we get Δ_f H°(Cl⁻, aq) = -167.08 kJ/mol (because the other three terms are zero).
  • By continuing with other reactions, a consistent set of ion formation enthalpies can be built.

🌡️ Temperature dependence

🌡️ Molar reaction heat capacity

Molar reaction heat capacity at constant pressure (Δ_r C_p): the rate at which the heat capacity C_p changes with advancement ξ at constant T and p.

  • Formula: Δ_r C_p = sum over all species i of (stoichiometric number_i times C_p,i)
  • Relates to temperature dependence: (∂Δ_r H / ∂T) at constant p and ξ = Δ_r C_p
  • Under standard state conditions: dΔ_r H° / dT = Δ_r C°_p

🌡️ Kirchhoff equation

For integral reaction enthalpy:

  • ΔH(rxn, T'') = ΔH(rxn, T') + integral from T' to T'' of ΔC_p dT
  • Where ΔC_p = C_p(ξ₂) - C_p(ξ₁) is the difference in heat capacities at final and initial advancement.
  • When ΔC_p is essentially constant: ΔH(rxn, T'') = ΔH(rxn, T') + ΔC_p (T'' - T')
  • Physical meaning: ΔC_p equals the difference in slopes of enthalpy vs. temperature for the system at final vs. initial advancement; the product ΔC_p × (temperature difference) equals the change in reaction enthalpy.

For molar differential reaction enthalpy:

  • Δ_r H(T'', ξ) = Δ_r H(T', ξ) + integral from T' to T'' of Δ_r C_p(T, ξ) dT
  • Analogous to the integral form, using molar differential quantities.

Example from the excerpt:

  • An exothermic reaction with negative ΔC_p results in a more negative ΔH(rxn) at higher temperature.
  • Don't confuse: the sign of ΔC_p determines whether the reaction becomes more or less exothermic/endothermic with increasing temperature.
69

Enthalpies of Solution and Dilution

11.4 Enthalpies of Solution and Dilution

🧭 Overview

🧠 One-sentence thesis

Enthalpies of solution and dilution quantify the enthalpy changes when solute dissolves into solvent or when solvent is added to an existing solution, both occurring in closed systems where material transfers between phases without changing total amounts.

📌 Key points (3–5)

  • Solution vs dilution processes: solution transfers solute from a pure phase into solvent/solution; dilution transfers solvent into an existing solution.
  • Closed-system nature: both processes occur in closed systems with initially two phases—total amounts of solvent and solute remain constant, but pure-phase amounts decrease as the process advances.
  • Molar differential enthalpy of solution: measures the rate of enthalpy change per amount of solute transferred at constant temperature and pressure, depending only on solution molality (not total amount).
  • Infinite dilution limit: at infinite dilution, the molar differential and integral enthalpies of solution are identical because the partial molar enthalpy becomes independent of composition.
  • Common confusion: don't confuse solution (adding solute to solvent) with dilution (adding solvent to solution)—both involve phase transfer but in opposite directions.

🔄 Solution and dilution processes

🧪 What is a solution process

Solution process: a solute is transferred from a pure solute phase (solid, liquid, or gas) to a solvent or solution phase.

  • The solute starts in a pure phase and moves into the liquid solution.
  • The system is closed: no material enters or leaves the overall system.
  • As the process advances (measured by ξ_sol), the amount of pure solute decreases while the amount in solution increases.
  • Example: solid salt crystals dissolving into water—the salt moves from the pure solid phase into the aqueous solution phase.

💧 What is a dilution process

Dilution process: solvent is transferred from a pure solvent phase to a solution phase.

  • The solvent starts pure and is added to an existing solution.
  • Again, the system is closed with two phases initially.
  • As the process advances (measured by ξ_dil), the amount of pure solvent decreases while the solution volume increases.
  • Example: pouring pure water into a salt solution—the water moves from the pure liquid phase into the solution phase.

🔍 How to distinguish solution from dilution

ProcessWhat transfersFrom → ToAdvancement variable
SolutionSolutePure solute phase → Solutionξ_sol
DilutionSolventPure solvent phase → Solutionξ_dil
  • Both are related processes but involve different materials moving.
  • Don't confuse: in solution, you're adding solute; in dilution, you're adding solvent.
  • The excerpt emphasizes that both take place in closed systems with at least two phases initially.

📐 Molar differential enthalpy of solution

📏 Definition and meaning

Molar differential enthalpy of solution (Δ_sol H): the rate of change of enthalpy H with the advancement ξ_sol at constant temperature and pressure, where ξ_sol is the amount of solute transferred.

  • Mathematically: Δ_sol H = (∂H/∂ξ_sol) at constant T, p, and amount of solvent n_A.
  • It measures "how much enthalpy changes per mole of solute dissolved."
  • The value depends only on the solution molality (concentration), not on the total amount of solution present.

🧮 How it relates to partial molar enthalpies

  • For the solution reaction written as B → B(sln), the general relation gives:
    • Δ_sol H = H_B − H*_B
    • H_B is the partial molar enthalpy of the solute in the solution.
    • H*_B is the molar enthalpy of the pure solute at the same T and p.
  • The difference captures the enthalpy change when one mole of solute moves from the pure state into the solution environment.

🌊 Infinite dilution limit

Molar enthalpy of solution at infinite dilution (Δ_sol H^∞): the rate of change of H with ξ_sol when the solute is transferred to a solution with the thermal properties of an infinitely dilute solution.

  • Think of it as "the enthalpy change per mole of solute transferred to a very large volume of pure solvent."
  • Formula: Δ_sol H^∞ = H^∞_B − H*_B
    • H^∞_B is the partial molar enthalpy of solute at infinite dilution.
  • Key property: because H^∞_B and H*_B are both independent of solution composition, the molar differential and integral enthalpies of solution at infinite dilution are the same.
  • Don't confuse: at finite concentrations, differential and integral quantities differ; at infinite dilution, they coincide because composition effects vanish.

🔧 General applicability

🔄 Extension to other state functions

  • The excerpt notes that the equations for enthalpies of solution and dilution apply to any extensive state function.
  • You can replace H by another extensive property (e.g., volume, Gibbs energy) to obtain analogous solution and dilution properties.
  • Example: replacing H with G would give molar differential Gibbs energy of solution, following the same mathematical structure.
70

Reaction Calorimetry

11.5 Reaction Calorimetry

🧭 Overview

🧠 One-sentence thesis

Reaction calorimetry determines the molar integral reaction enthalpy by measuring temperature changes in controlled vessels, with constant-pressure and bomb calorimeters using different correction paths to convert experimental temperature data into standard-state enthalpy values.

📌 Key points (3–5)

  • What is measured vs. what is wanted: calorimeters measure temperature change, but the goal is the molar integral reaction enthalpy ΔH_m(rxn) at constant temperature and pressure.
  • Two main types: constant-pressure reaction calorimeters (for liquid-phase processes) and bomb calorimeters (for combustion reactions at constant volume).
  • Key distinction: constant-pressure calorimeters are open to atmosphere and unsuitable for gases; bomb calorimeters are sealed high-pressure vessels designed for complete combustion.
  • Common confusion: the experimental enthalpy change ΔH(expt) is not the reaction enthalpy ΔH(rxn, T₁) because temperature changes during the experiment; corrections using energy equivalents are needed.
  • Why it matters: bomb calorimetry provides standard molar enthalpies of combustion, which via Hess's law yield formation enthalpies for calculating reaction enthalpies of countless reactions.

🔬 General principles of reaction calorimetry

🎯 What reaction calorimetry measures

Reaction calorimetry is used to evaluate the molar integral reaction enthalpy ΔH_m(rxn) of a reaction or other chemical process at constant temperature and pressure.

  • The actual measurement is a temperature change in the calorimeter.
  • The desired quantity is the enthalpy change at constant temperature.
  • This mismatch requires corrections.

🏗️ Common calorimeter design

  • A reaction vessel surrounded by an outer jacket.
  • The jacket may be:
    • Adiabatic type: minimizes heat transfer to surroundings.
    • Isothermal-jacket type: maintains constant jacket temperature.
  • A temperature-measuring device is immersed in the vessel or in thermal contact with it.
  • Key difference from heat capacity calorimeters: work is kept deliberately small to minimize internal energy and enthalpy changes during the experiment.

🧪 Constant-pressure reaction calorimeter

🌊 Suitable processes

  • Contents are usually open to the atmosphere → unsuitable for processes involving gases.
  • Convenient for:
    • Liquid-phase chemical reactions.
    • Dissolution of solid or liquid solute in liquid solvent.
    • Dilution of a solution with solvent.

📈 Experimental procedure

  • Initiation: reactants are brought into contact.
  • Observation: temperature is measured over time, starting before initiation and ending after advancement ξ reaches a final value with no further change.
  • Heating/cooling curve: resembles those in heat capacity measurements (Figs. 7.3 and 7.4).
  • Two key points:
    • Temperature T₁ before reaction initiation.
    • Temperature T₂ after ξ reaches its final value.

🔄 Path analysis and corrections

The excerpt describes three paths at constant pressure (Fig. 11.11):

PathDescriptionSymbol
Experimental processReactants at T₁ → products at T₂ΔH(expt)
Reaction at constant T₁Reactants → products at T₁ΔH(rxn, T₁)
Temperature change of productsProducts at T₁ → products at T₂ΔH(P)
  • The paths have the same initial state but different final states.
  • Key relation: ΔH(expt) = ΔH(rxn, T₁) + ΔH(P).
  • ΔH(P) = ε_P(T₂ − T₁), where ε_P is the energy equivalent (average heat capacity) when the calorimeter contains products.

🧮 Calculating the reaction enthalpy

From the path independence of enthalpy:

ΔH(rxn, T₁) = ε_P(T₂ − T₁) + ΔH(expt)

  • ΔH(expt) is evaluated using methods from Sec. 7.3.2:
    • Adiabatic calorimeter: ΔH(expt) = ε(t₂ − t₁), where ε is the energy equivalent and t₁, t₂ are times at T₁, T₂.
    • Isothermal-jacket calorimeter: use Eq. 7.3.28 with electrical work set to zero.
  • ε_P is measured in a second experiment with electric heating.

Molar integral reaction enthalpy:

ΔH_m(rxn) = ΔH(rxn, T₁)/Δξ = [ε_P(T₂ − T₁) + ΔH(expt)]/Δξ

  • ΔH(expt) is small, so ΔH_m(rxn) ≈ ε_P(T₂ − T₁)/Δξ.
  • If T₂ > T₁ (exothermic), ΔH_m(rxn) is negative (heat must leave to return to T₁).
  • If T₂ < T₁ (endothermic), ΔH_m(rxn) is positive.

⚠️ Important distinction

  • ΔH_m(rxn) is not the same as the molar differential reaction enthalpy Δ_r H = (∂H/∂ξ)_{T,p} unless phases can be treated as ideal mixtures.
  • Small corrections are needed to obtain the standard molar reaction enthalpy Δ_r H° from ΔH_m(rxn).

💣 Bomb calorimeter (constant-volume)

🔥 Typical use

  • Complete combustion of a solid or liquid substance in excess oxygen.
  • Combustion initiated with electrical ignition.
  • May include side reactions (e.g., nitrogen oxide formation) and auxiliary reactions.

🎯 Goal

Evaluate the standard molar enthalpy of combustion Δ_c H° at a reference temperature T_ref (often 298.15 K = 25.00 °C).

  • Precision can be ~0.01% with careful work and detailed corrections.

🏗️ Apparatus design (Fig. 11.12)

  • Bomb vessel: thick-walled cylindrical metal vessel ("bomb" because it withstands high pressure).
  • Sealed with gas-tight screw cap.
  • During reaction, sealed bomb is immersed in water in the calorimeter, surrounded by a jacket.
  • System: everything inside the jacket (calorimeter walls, water, bomb vessel, and bomb contents).

🧪 Experimental setup

  1. Weighed sample placed in metal sample holder.
    • Volatile liquids encapsulated in thin glass bulb or confined with cellulose tape.
  2. If combustion produces H₂O, a small known mass of liquid water is placed in the bomb to saturate the gas space.
  3. Oxygen gas admitted to ~30 bar total pressure.
  4. Sealed bomb immersed in known mass of water.
  5. Precision thermometer and stirrer immersed in water.

📊 Experimental procedure

  • Initial time t₁: temperature T₁ (uniform throughout system) after slow, constant drift is observed.
  • Ignition: circuit closed at or soon after t₁.
  • Temperature rise: if exothermic, temperature rapidly increases over several minutes.
  • Final time t₂: temperature T₂ (uniform again) after slow, constant drift resumes.
  • Pressure behavior: water outside bomb stays at atmospheric pressure; inside the bomb, pressure is not constant (volume is essentially constant).

🔄 Path analysis for bomb calorimeter

Three paths at constant volume (Fig. 11.13):

PathDescriptionSymbol
Experimental processReactants at T₁ → products at T₂ΔU(expt)
Temperature change of reactantsReactants at T₁ → reactants at T₂ΔU(R)
Isothermal bomb processReactants → products at T₂ΔU(IBP, T₂)
  • Key relation: ΔU(expt) = ΔU(R) + ΔU(IBP, T₂).
  • ΔU(R) = ε_R(T₂ − T₁), where ε_R is the energy equivalent (average heat capacity) with reactants present.
  • ε_R is obtained in a separate calibration experiment, often using benzoic acid (whose combustion internal energy is precisely known).

🧮 Five main steps to evaluate Δ_c H°

🔢 Step 1: Determine ΔU(IBP, T₂)

The isothermal bomb process is the idealized process that would have occurred if the reaction or reactions had taken place in the calorimeter at constant temperature.

From the path relation:

ΔU(IBP, T₂) = ε_R(T₂ − T₁) + ΔU(expt)

ΔU(expt) is small and includes:

  1. Electrical work w_ign from ignition circuit (known).
  2. Heat transfer (minimized but not eliminated by jacket).
  3. Mechanical stirring work.
  4. Electrical work from thermometer.

ΔU(IBP, T₂) = ε_R(T₂ − T₁) + w_ign + ΔU'(expt)

  • ΔU'(expt) is the internal energy change due to heat, stirring, and temperature measurement.
  • Evaluated using energy equivalent and observed temperature drift rates at t₁ and t₂.

🌡️ Step 2: Correct to reference temperature T_ref

If T₂ and T_ref are close, use a modified Kirchhoff equation:

ΔU(IBP, T_ref) = ΔU(IBP, T₂) + [C_V(P) − C_V(R)](T_ref − T₂)

  • C_V(P) and C_V(R) are heat capacities at constant volume of bomb contents with products and reactants, respectively.

🎯 Step 3: Reduction to standard states

Calculate the standard internal energy change for the main combustion reaction at T_ref.

Hypothetical three-step process at T_ref:

  1. Each substance changes from standard state to initial state of isothermal bomb process.
  2. Isothermal bomb process occurs (main + side + auxiliary reactions).
  3. Each substance changes from final state to standard state.

ΔU°(cmb, T_ref) = ΔU(IBP, T_ref) + (Washburn corrections) − Σ_i Δξ_i Δ_r U°(i)

  • Sum over i is for side and auxiliary reactions.
  • Washburn corrections: internal energy changes for hypothetical physical processes at T_ref (substances changing between actual states and standard states).

📐 Step 4: Calculate standard molar internal energy of combustion

Δ_c U°(T_ref) = ΔU°(cmb, T_ref)/Δξ_c

  • Δξ_c is the advancement of the main combustion reaction.

🔄 Step 5: Convert to standard molar enthalpy of combustion

From H_i = U_i + pV_i:

Δ_c H°(T_ref) = Δ_c U°(T_ref) + p° Σ_i ν_i V°_i

Since condensed-phase molar volumes are much smaller than gas volumes:

Δ_c H°(T_ref) ≈ Δ_c U°(T_ref) + p° Σ_{i,g} ν_i V°_i(g)

For ideal gases in standard state, V°_i = RT/p°, so:

Δ_c H°(T_ref) = Δ_c U°(T_ref) + Σ_{i,g} ν_i RT_ref

  • Sum includes only gaseous reactants and products.

🔧 Washburn corrections details

Example: complete combustion of a C, H, O compound → CO₂ and H₂O (no side reactions).

Initial state of isothermal bomb process:

  • Pure reactant.
  • Liquid water with dissolved O₂.
  • Gaseous mixture of O₂ and H₂O.
  • All at high pressure p₁.

Final state:

  • Liquid water with dissolved O₂ and CO₂.
  • Gaseous mixture of O₂, H₂O, and CO₂.
  • All at pressure p₂.

Standard states chosen:

  • Reactant compound: pure solid or liquid at p° = 1 bar.
  • H₂O: pure liquid at p°.
  • O₂ and CO₂: pure ideal gases at p°.

Don't confuse: A single standard state for each substance must be used in Eq. 11.5.9 to correctly give the standard internal energy of combustion.

Amounts of each substance in each phase (initial and final states) are calculated from:

  • Internal volume of bomb vessel.
  • Mass of reactant.
  • Initial amount of H₂O.
  • Initial O₂ pressure.
  • Water vapor pressure.
  • Solubilities of O₂ and CO₂ (estimated from Henry's law constants).
  • Stoichiometry of combustion reaction.

🔬 Other calorimeter types

🧊 Phase-change calorimeter

  • Two coexisting phases of a pure substance in thermal contact with reaction vessel and adiabatic jacket.
  • At constant pressure, the univariant subsystem is at the fixed temperature of the phase transition.
  • Thermal energy from reaction is transferred isothermally to/from coexisting phases.
  • Measured by volume change of the phase transition, not temperature change.
  • Limitation: can only measure at the phase-transition temperature.
  • Examples:
    • Bunsen ice calorimeter: ice–water transition at 0 °C.
    • Diphenyl ether: solid–liquid transition at 26.9 °C (large volume change).
  • Especially useful for slow reactions.

🌡️ Heat-flow calorimeter

  • Variation of isothermal-jacket calorimeter.
  • Uses a thermopile to continuously measure temperature difference between reaction vessel and constant-temperature heat sink (outer jacket).
  • Heat transfer mostly through thermocouple wires.
  • Heat transfer is proportional to the temperature difference integrated over time.
  • Best method for extremely slow reactions; also usable for rapid reactions.

🔥 Flame calorimeter

  • Flow system: gaseous oxidant (O₂, F₂, etc.) reacts with gaseous fuel.
  • Heat transfer between flow tube and heat sink measured with thermopile.
  • Similar principle to heat-flow calorimeter but for flowing gases.

🌐 Importance of bomb calorimetry

📚 Foundation for thermochemical data

  • Bomb calorimetry is the principal means to evaluate standard molar enthalpies of combustion of elements and compounds.
  • From combustion enthalpies, using Hess's law, standard molar enthalpies of formation of compounds are calculated.
  • From formation values of only a few compounds, standard molar reaction enthalpies of innumerable reactions can be calculated with Hess's law.

Example workflow:

  1. Measure Δ_c H° for compound via bomb calorimetry.
  2. Calculate Δ_f H° for compound using Hess's law and known Δ_f H° of combustion products.
  3. Use Δ_f H° values to calculate Δ_r H° for any reaction involving that compound.

This hierarchical approach makes bomb calorimetry data extremely valuable for thermochemistry.

71

Adiabatic Flame Temperature

11.6 Adiabatic Flame Temperature

🧭 Overview

🧠 One-sentence thesis

The adiabatic flame temperature can be estimated by assuming a constant-pressure, adiabatic combustion process in an ideal-gas mixture and solving for the temperature at which the standard enthalpy of combustion balances the enthalpy change of heating the products.

📌 Key points (3–5)

  • What it estimates: the temperature of a flame formed when oxygen or air reacts with a fuel in a flowing gas mixture.
  • Core assumption: the combustion occurs adiabatically at constant pressure (standard pressure), with no heat transfer to surroundings.
  • How to calculate: set the sum of the standard molar enthalpy of combustion and the enthalpy change from heating the products equal to zero.
  • Why actual flames are cooler: the process is never completely adiabatic, and high temperatures cause product dissociation and side reactions.
  • Common confusion: the calculated temperature is an estimate under idealized conditions, not the true flame temperature.

🔥 What the adiabatic flame temperature represents

🔥 The physical scenario

  • A segment of a flowing gas mixture (oxygen or air + fuel) is treated as a closed system.
  • As combustion takes place, the temperature of this segment increases.
  • The goal is to estimate the final temperature when the reaction reaches equilibrium.

🧪 Key simplifying assumptions

The calculation relies on several approximations:

  • The reaction occurs at constant pressure equal to the standard pressure (1 bar).
  • The process is adiabatic (no heat exchange with surroundings).
  • The gas behaves as an ideal-gas mixture.

Example: Imagine a small "packet" of fuel-air mixture moving through a burner; we assume it burns without losing heat to the walls or surroundings.

🧮 The calculation method

🧮 The governing equation

The principle is similar to that used for a constant-pressure calorimeter. The equation is:

Equation 11.6.1: (delta)(xi) times (delta_c H°(T₁)) + integral from T₁ to T₂ of C_p(P) dT = 0

Where:

  • (delta)(xi) is the change in reaction advancement (how far the reaction proceeds).
  • (delta_c H°(T₁)) is the standard molar enthalpy of combustion at the initial temperature T₁.
  • C_p(P) is the heat capacity at constant pressure of the product mixture.
  • T₂ is the final (flame) temperature we want to find.

🔍 Why this equation works

  • Because the reaction is assumed to be adiabatic at constant pressure, the experimental enthalpy change (delta H_expt) is zero.
  • This means the enthalpy released by combustion must exactly balance the enthalpy needed to heat the products from T₁ to T₂.
  • The sum of the reaction enthalpy and the heating enthalpy must equal zero.

📐 Connection to calorimeter paths

The excerpt references Figure 11.11 (from page 335), which shows paths used for constant-pressure calorimeters:

  • Path 1: Combustion reaction at initial temperature T₁ releases enthalpy.
  • Path 2: Products are heated from T₁ to T₂, absorbing enthalpy.
  • For an adiabatic process, these two enthalpy changes cancel out.

⚠️ Why actual flames are cooler

⚠️ Limitations of the estimate

The value of T₂ that satisfies the equation is the estimated flame temperature, not the actual temperature. Several factors cause real flames to be cooler:

FactorEffect
Never completely adiabaticSome heat is always lost to surroundings (radiation, conduction through walls)
Product dissociationAt high temperatures, combustion products may break apart, absorbing energy
Side reactionsAdditional reactions beyond the main combustion consume or release energy

🔄 Don't confuse with

  • Ideal vs. real: The calculation gives an upper bound under ideal conditions; actual measurements will always be lower.
  • Equilibrium assumption: The method assumes the reaction reaches equilibrium, but in fast-flowing systems this may not fully occur.

🔗 Context and application

🔗 Relation to other calorimetry methods

The excerpt places adiabatic flame temperature in the context of other calorimeters:

  • Phase-change calorimeters: Use coexisting phases at fixed temperature (e.g., ice-water at 0°C).
  • Heat-flow calorimeters: Use thermopiles to measure temperature differences continuously.
  • Flame calorimeters: Flow systems where oxidant reacts with gaseous fuel; heat transfer measured with thermopiles.

The adiabatic flame temperature method differs because it assumes no heat transfer (adiabatic), whereas flame calorimeters measure the heat transfer.

📝 Practical use

  • Problem 11.9 (mentioned in the excerpt) presents an application of this calculation.
  • The method is useful for quick estimates of flame temperatures in combustion engineering and safety analysis.
72

Gibbs Energy and Reaction Equilibrium

11.7 Gibbs Energy and Reaction Equilibrium

🧭 Overview

🧠 One-sentence thesis

The molar reaction Gibbs energy determines the direction of spontaneous chemical change, and reaction equilibrium is reached universally when this quantity becomes zero, regardless of the path taken.

📌 Key points (3–5)

  • What drives spontaneity: At constant T and p, reactions proceed in the direction that decreases Gibbs energy G; the molar reaction Gibbs energy Δ_r G (the rate of change of G with advancement ξ) determines whether ξ increases or decreases spontaneously.
  • Universal equilibrium criterion: Reaction equilibrium occurs when Δ_r G = Σ_i ν_i μ_i = 0, a condition that holds regardless of whether T and p are constant during the approach to equilibrium.
  • Pure phases vs. mixtures: In systems of pure phases, Δ_r G is constant at fixed T and p, so reactions go to completion unless Δ_r G = 0; in systems containing mixtures, G has a minimum at equilibrium due to the mixing term.
  • Common confusion: The minimum of G at constant T and p corresponds to reaction equilibrium for reactions involving mixtures, but for reactions at constant T and V (where p changes), equilibrium is at the minimum of Helmholtz energy A, not G.
  • Why it matters: The standard molar reaction Gibbs energy Δ_r G° determines the equilibrium position—more negative Δ_r G° shifts equilibrium toward products, more positive Δ_r G° shifts it toward reactants.

🔬 Defining the molar reaction Gibbs energy

🔬 What Δ_r G measures

Molar reaction Gibbs energy (Δ_r G): the rate at which the Gibbs energy G of a closed system changes with respect to the advancement ξ at constant temperature T and pressure p.

  • Defined as Δ_r G = Σ_i ν_i μ_i, where ν_i is the stoichiometric coefficient and μ_i is the chemical potential of species i.
  • Also given by the partial derivative: Δ_r G = (∂G/∂ξ)_{T,p} for a closed system.
  • The total differential of G for a closed system is: dG = −S dT + V dp + Δ_r G dξ.

🔍 How Δ_r G relates to advancement

  • Δ_r G tells you whether the advancement ξ will spontaneously increase or decrease.
  • If Δ_r G is positive, ξ spontaneously decreases (reaction shifts left).
  • If Δ_r G is negative, ξ spontaneously increases (reaction shifts right).
  • During a spontaneous process, dξ and Δ_r G have opposite signs, so the term Δ_r G dξ is negative, ensuring dG < −S dT + V dp (the spontaneity criterion).

⚖️ The universal equilibrium condition

⚖️ Equilibrium criterion: Δ_r G = 0

Reaction equilibrium condition: Δ_r G = Σ_i ν_i μ_i = 0.

  • This condition is independent of whether equilibrium is approached at constant T and p or by some other path.
  • It is a universal criterion valid for any reaction in any closed system.
  • At equilibrium, G has its minimum value for the given T and p (if the system contains at least one mixture phase and equilibrium can be approached from either direction).

🧩 Why this condition is universal

  • The excerpt derives this condition rigorously by considering an isolated system with multiple phases and a chemical reaction.
  • At equilibrium:
    • All phases have the same temperature.
    • All phases have the same pressure.
    • Each species has the same chemical potential in all phases where it is present.
    • The molar reaction Gibbs energy of each phase is zero.
  • The condition Δ_r G = 0 emerges from the requirement that entropy S is maximized in an isolated system, which translates to G being minimized at constant T and p.

🔄 How Δ_r G changes during spontaneous processes

  • If Δ_r G is positive, spontaneous change decreases Δ_r G (by decreasing ξ).
  • If Δ_r G is negative, spontaneous change increases Δ_r G (by increasing ξ).
  • When the system reaches equilibrium, Δ_r G becomes zero, regardless of the path taken.

🧪 Reactions in pure phases vs. mixtures

🧪 Pure phases: constant Δ_r G

  • When each reactant and product is in a separate pure phase (e.g., CaCO₃(s) → CaO(s) + CO₂(g)), the chemical potential of each substance remains constant at constant T and p.
  • Therefore, Δ_r G is constant and depends only on T and p.
  • Three cases:
    • Δ_r G < 0: Reaction proceeds spontaneously to the right until a reactant is exhausted ("goes to completion").
    • Δ_r G > 0: Reaction proceeds spontaneously to the left until a product is exhausted.
    • Δ_r G = 0: Reactants and products can remain in equilibrium at all values of ξ.
  • The plot of G versus ξ is a straight line with slope Δ_r G.
  • Don't confuse: This behavior is similar to phase transitions of a pure substance—only one phase is present at equilibrium unless the transition Gibbs energy is zero.

🌀 Mixtures: G has a minimum

  • If any reactant or product is a constituent of a mixture, the plot of G versus ξ (at constant T and p) exhibits a minimum with slope zero.
  • At constant T and p, ξ changes spontaneously toward decreasing G until the minimum is reached, where Δ_r G = 0 and the system is at equilibrium.
  • Example: For the reaction A(g) → B(g) in an ideal gas mixture, the Gibbs energy is:
    • G(ξ) − G(0) = ξ Δ_r G° + RT Σ_i n_i ln y_i − RT Σ_i n_{i,0} ln y_{i,0} + RT (Σ_i ν_i) ξ ln(p/p°).
  • The mixing term RT Σ_i n_i ln y_i is responsible for the minimum; it is entirely an entropy effect (equal to −nT ΔS_m^id(mix)).

🎯 How Δ_r G° affects equilibrium position

  • The standard molar reaction Gibbs energy Δ_r G° = Σ_i ν_i μ_i° depends only on T and the choice of standard states.
  • More negative Δ_r G° → equilibrium position closer to the product side.
  • More positive Δ_r G° → equilibrium position closer to the reactant side.
  • Δ_r G° = 0 → equilibrium is roughly in the middle.

🔧 Equilibrium at constant T and V

🔧 Using Helmholtz energy A instead of G

  • For a reaction at constant T and volume V (where pressure may change), the relevant thermodynamic potential is the Helmholtz energy A, not G.
  • The total differential of A in a closed system with a reaction is:
    • dA = −S dT − p dV + (Σ_i ν_i μ_i) dξ.
  • The coefficient of dξ is Σ_i ν_i μ_i = (∂A/∂ξ)_{T,V}.
  • At constant T and V, the reaction reaches equilibrium when (∂A/∂ξ)_{T,V} = 0, which is the same condition Σ_i ν_i μ_i = 0.
  • Don't confuse: At constant T and V, equilibrium is at the minimum of A, not the minimum of G (though the condition Δ_r G = 0 still holds).

🔄 Example: A → 2B in an ideal gas

  • As this reaction proceeds to the right at constant T:
    • Volume increases if pressure is held constant.
    • Pressure increases if volume is held constant.
  • At constant T and p, the minimum of G is at the equilibrium position.
  • At constant T and V, the minimum of A is at the equilibrium position (which is at a different value of ξ than the constant-p case).
  • The condition Δ_r G = 0 is satisfied along the curve of equilibrium states, not necessarily at the minimum of G when V is constant.

📐 Standard molar reaction Gibbs energy

📐 Definition and properties

Standard molar reaction Gibbs energy (Δ_r G°): Δ_r G° = Σ_i ν_i μ_i°, the sum of stoichiometric coefficients times standard chemical potentials.

  • Depends only on temperature T and the choice of standard states for each reactant and product.
  • Does not depend on the actual composition, pressure, or electric potential of the system.

⚡ Relation to activities

  • The general relation between Δ_r G and Δ_r G° is:
    • Δ_r G = Δ_r G° + RT Σ_i ν_i ln a_i, where a_i is the activity of species i.
  • This relation holds when any ions are all in the same phase or in phases of equal electric potential (because the term involving electric potential Φ sums to zero due to charge conservation).
  • For a reaction at a given temperature with ions in the same phase, Δ_r G depends only on the activities of reactants and products, not on the electric potential of the phase.

🔑 Why Δ_r G° matters

  • Δ_r G° is the Gibbs energy change for the reaction of pure reactants to form pure products under standard-state conditions.
  • It sets the "baseline" for the equilibrium position.
  • The actual value of Δ_r G at any state is adjusted from Δ_r G° by the mixing term (RT Σ_i ν_i ln a_i) and any pressure correction.
73

The Thermodynamic Equilibrium Constant

11.8 The Thermodynamic Equilibrium Constant

🧭 Overview

🧠 One-sentence thesis

The thermodynamic equilibrium constant K, defined as the proper quotient of activities at equilibrium, depends only on temperature and the choice of standard states—not on pressure or composition—and is directly related to the standard molar reaction Gibbs energy through the equation ΔrG° = –RT ln K.

📌 Key points (3–5)

  • What K represents: the value of the reaction quotient Q_rxn (proper quotient of activities) when the system reaches reaction equilibrium.
  • K depends only on T and standard states: no other condition (pressure, composition, volume) can affect K's value for a given reaction.
  • Relationship to ΔrG°: the standard molar reaction Gibbs energy determines K via K = exp(–ΔrG°/RT); negative ΔrG° means K > 1, positive ΔrG° means K < 1.
  • Common confusion: K (thermodynamic equilibrium constant) vs K_p (equilibrium constant on a pressure basis)—K_p can vary with pressure at constant temperature, but K cannot.
  • Practical implication: very large K means reactants are nearly exhausted at equilibrium; very small K means products have very small activities at equilibrium.

🧮 Defining the reaction quotient and K

🧮 The reaction quotient Q_rxn

Reaction quotient (activity quotient), Q_rxn: the product of activities of all species, each raised to the power of its stoichiometric number νᵢ.

  • Formula in words: Q_rxn equals the product over all species i of (activity of i) raised to the power νᵢ.
  • Because νᵢ is positive for products and negative for reactants, Q_rxn is a quotient with products in the numerator and reactants in the denominator.
  • Each activity is raised to a power equal to the stoichiometric coefficient in the reaction equation—this is called a proper quotient.
  • Q_rxn is dimensionless and changes as the reaction advances (it depends on T, p, and composition).

Example: For N₂(g) + 3 H₂(g) → 2 NH₃(g), Q_rxn = (a_NH₃)² / [(a_N₂)(a_H₂)³].

🎯 The thermodynamic equilibrium constant K

Thermodynamic equilibrium constant, K: the value of Q_rxn under equilibrium conditions; the product of activities (each raised to its stoichiometric number) evaluated at equilibrium.

  • Formula in words: K equals the product over all species i of (activity of i at equilibrium) raised to the power νᵢ.
  • K is dimensionless.
  • The excerpt notes that IUPAC sometimes uses the symbol K° and the name "standard equilibrium constant," but warns this can cause confusion—K refers to an equilibrium state, not to reactants and products in their standard states.

🔗 Connecting Q_rxn, K, and ΔrG

The molar reaction Gibbs energy is:

ΔrG = ΔrG° + RT ln Q_rxn

  • At equilibrium, ΔrG = 0 and Q_rxn = K, so:

ΔrG° = –RT ln K

or equivalently:

K = exp(–ΔrG°/RT)

  • This shows that K is determined entirely by ΔrG° (the standard molar reaction Gibbs energy), which depends only on T and the choice of standard states.
  • If ΔrG° is positive, K < 1; if ΔrG° is negative, K > 1.

🔬 Standard molar reaction Gibbs energy ΔrG°

🔬 Definition and properties

Standard molar reaction Gibbs energy, ΔrG°: the sum over all species i of (stoichiometric number νᵢ) × (standard chemical potential μ°ᵢ).

  • Formula in words: ΔrG° equals the sum over all species i of νᵢ μ°ᵢ.
  • Because each μ°ᵢ is a function only of T, ΔrG° depends only on T and the choice of standard states for each reactant and product.
  • The excerpt derives ΔrG from the general relation for chemical potential (involving activity and electric potential) and shows that for reactions where all ions are at the same electric potential, ΔrG depends only on activities and is independent of the actual electric potential values.

📐 Calculating ΔrG° from formation data

The excerpt provides the relation:

ΔrG° = ΔrH° – T ΔrS°

where:

  • ΔrH° is the standard molar reaction enthalpy.
  • ΔrS° is the standard molar reaction entropy (sum over all species i of νᵢ S°ᵢ).

For a formation reaction of a substance:

ΔfG° = ΔfH° – T (sum over reactants and product of νᵢ S°ᵢ)

  • Tabulated values of ΔfG° (standard molar Gibbs energy of formation) are available; Appendix H gives values at 298.15 K.
  • For any reaction, ΔrG° can be calculated from formation data using (analogous to Hess's law):

ΔrG° = sum over all species i of νᵢ ΔfG°(i)

  • The excerpt notes that ΔfH° values come from calorimetry, and absolute molar entropy values S°ᵢ come from heat capacity data or statistical mechanics—so K can be evaluated entirely from calorimetric measurements, a goal of early 20th-century thermodynamics.

⚡ Special conventions for aqueous ions

  • For ions in aqueous solution, the reference values are S°_m = 0 and ΔfG° = 0 for H⁺(aq) at all temperatures (similar to the convention for ΔfH°).
  • These conventions correctly give ΔrS° and ΔrG° for reactions involving aqueous ions.
  • The relation ΔfG° = ΔfH° – T ΔrS° does not apply directly to a single ion (because you cannot write a formation reaction for a single ion).
  • Instead, the excerpt provides formulas for cations and anions involving hypothetical reactions with H⁺(aq) and H₂(g):
    • For a cation M^(z⁺): ΔfG°(M^(z⁺)) = ΔfH°(M^(z⁺)) – T[S°_m(M^(z⁺)) – (sum over elements of S°ᵢ) + (z/2)S°_m(H₂)]
    • For an anion X^(z⁻): ΔfG°(X^(z⁻)) = ΔfH°(X^(z⁻)) – T[S°_m(X^(z⁻)) – (sum over elements of S°ᵢ) – (|z|/2)S°_m(H₂)]

⚙️ K for gas-phase reactions

⚙️ Activities and fugacity coefficients

For a reaction in a gaseous mixture, the standard state of each species is the pure gas behaving ideally at standard pressure p°.

  • Each activity is given by: a_i(g) = f_i / p° = γ_i p_i / p°, where γ_i is the fugacity coefficient.
  • Substituting into the definition of K gives:

K = [product over all i of (γ_i)^νᵢ at equilibrium] × [product over all i of (p_i)^νᵢ at equilibrium] × [(p°)^(–sum of νᵢ)]

  • The first factor is the proper quotient of fugacity coefficients at equilibrium.
  • The second factor is the proper quotient of equilibrium partial pressures, called K_p.
  • The third factor makes K dimensionless.

🔄 K_p vs K

Equilibrium constant on a pressure basis, K_p: the proper quotient of equilibrium partial pressures.

  • Formula in words: K_p equals the product over all species i of (partial pressure of i at equilibrium) raised to the power νᵢ.
  • K_p is dimensionless only if the sum of stoichiometric numbers equals zero.
  • Don't confuse K and K_p: K_p can vary at constant temperature (e.g., when you compress an ideal gas mixture isothermally, fugacity coefficients change from 1, the mixture becomes nonideal, and K_p must change to keep K constant). K_p is not a thermodynamic equilibrium constant.

Example: For N₂(g) + 3 H₂(g) → 2 NH₃(g), the sum of νᵢ is –2, so:

  • K = [γ²_NH₃ / (γ_N₂ γ³_H₂)] × K_p × (p°)²
  • K_p = (p_NH₃)² / [(p_N₂)(p_H₂)³] at equilibrium.

💧 K for reactions in solution

💧 Different solute standard states

For solutes, the value of K depends on the choice of standard state (mole fraction, concentration, or molality basis).

  • At infinite dilution and standard pressure, each solute activity coefficient is unity and pressure factors are unity, so:
    • a_x,B = x_B (mole fraction basis)
    • a_c,B = c_B / c° (concentration basis)
    • a_m,B = m_B / m° (molality basis)
  • At infinite dilution, composition variables are related by: x_B = V_A c_B = M_A m_B, where V_A is the molar volume of pure solvent A and M_A is its molar mass.

🔀 Converting between K values on different bases

Combining the relations above, the activities at infinite dilution and pressure p° are related by:

  • a_x,B = V*_A c° a_c,B = M_A m° a_m,B

Because K is the product of activities raised to stoichiometric numbers, the values of K on different bases are related by:

K(x basis) = [product over all solutes B of (V*_A c°)^νB] × K(c basis)

K(c basis) = [product over all solutes B of (M_A m°)^νB] × K(m basis)

  • These relations are derived for infinite dilution at p°, but the excerpt emphasizes that for a given reaction at a given temperature, K is not affected by pressure or dilution, so these relations are valid under all conditions.

🎲 Physical meaning of K values

🎲 Extreme values of K

Activities typically cannot be many orders of magnitude greater than 1:

  • A partial pressure cannot exceed total pressure, so at 10 bar, a gas activity cannot exceed about 10.
  • A solute molarity is rarely much greater than 10 mol/dm³, corresponding to an activity of about 10.
  • Activities can, however, be extremely small.

Very large K:

  • At equilibrium, at least one reactant must have very small activity.
  • The reaction goes practically to completion; a limiting reactant is essentially exhausted.

Very small K:

  • At equilibrium, one or more products must have very small activity.
  • The reaction barely proceeds.

These are the two extremes of the trends shown in Figure 11.16 (page 349).

🔍 Equilibrium condition

At a fixed temperature, reaction equilibrium is attained if and only if the value of Q_rxn becomes equal to the value of K at that temperature.

  • K is the proper quotient of activities of species in reaction equilibrium.
  • The excerpt notes that Equation 11.8.10 (ΔrG° = –RT ln K) correctly relates ΔrG° and K only if they are both calculated with the same standard states.

Example: If you base the standard state of a solute on molality when calculating ΔrG°, the activity of that solute in the expression for K must also be based on molality.

📊 Evaluating K from thermodynamic data

📊 The calorimetric route

The relation K = exp(–ΔrG°/RT) allows you to evaluate K at a given temperature from ΔrG° at that temperature.

Method:

  1. Calculate ΔrG° from standard molar Gibbs energies of formation (ΔfG°) of each reactant and product using: ΔrG° = sum over all species i of νᵢ ΔfG°(i).
  2. Use ΔfG° = ΔfH° – T (sum of νᵢ S°ᵢ) for the formation reaction.
  3. Standard molar enthalpies of formation (ΔfH°) come from calorimetry.
  4. Absolute molar entropy values (S°ᵢ) come from heat capacity data or statistical mechanics.
  5. Calculate K from K = exp(–ΔrG°/RT).
  • The excerpt emphasizes that it is entirely feasible to use nothing but calorimetry to evaluate an equilibrium constant, a goal sought by thermodynamicists in the first half of the 20th century.
  • An alternative method (for reactions that can be carried out reversibly in a galvanic cell) is mentioned but deferred to Section 14.3.3.

📋 Using tabulated data

  • Extensive tables of ΔfG° values are available; Appendix H gives an abbreviated version at 298.15 K.
  • For aqueous ions, values are based on the reference convention ΔfG° = 0 and S°_m = 0 for H⁺(aq) at all temperatures.
  • Some ions have negative S°_m values in the tables; this simply means their standard molar entropies are less than that of H⁺(aq).

🌡️ Effects of temperature and pressure on equilibrium position

🌡️ Equilibrium shifts

The advancement ξ of a reaction in a closed system at constant T and p changes spontaneously until the system reaches reaction equilibrium at the minimum of G, where ξ = ξ_eq.

  • The value of ξ_eq depends in general on T and p.
  • When you change T or p of a closed system at equilibrium, ξ_eq usually changes and the reaction spontaneously shifts to a new equilibrium position.

🧮 Differential relations

The excerpt derives reciprocity relations from the total differential of G with T, p, and ξ as independent variables:

  • (∂ΔrG/∂T) at constant p and ξ = –ΔrS
  • (∂ΔrG/∂p) at constant T and ξ = ΔrV

where ΔrS and ΔrV are molar differential reaction quantities (molar reaction entropy and molar reaction volume).

The total differential of ΔrG is:

dΔrG = –ΔrS dT + ΔrV dp + (∂ΔrG/∂ξ) at constant T,p dξ

  • The coefficient (∂ΔrG/∂ξ) at constant T and p is the second derivative of G with respect to ξ.
  • At the minimum of G (reaction equilibrium), the slope ΔrG is zero and the second derivative is positive.
  • At equilibrium, ΔrG = 0, so ΔrS = ΔrH/T (from ΔrG = ΔrH – T ΔrS).

The excerpt sets up these relations but does not complete the analysis of how T and p shifts affect ξ_eq (this is deferred to Section 11.9).

Budget: 1000000 tokens remaining

74

Effects of Temperature and Pressure on Equilibrium Position

11.9 Effects of Temperature and Pressure on Equilibrium Position

🧭 Overview

🧠 One-sentence thesis

When temperature or pressure changes in a closed system at equilibrium, the equilibrium position shifts in a direction that opposes the change, with the shift direction determined by the signs of the reaction enthalpy and volume changes.

📌 Key points (3–5)

  • What determines equilibrium position: the advancement ξ_eq depends on temperature T and pressure p; changing T or p causes the equilibrium to shift to a new position.
  • Temperature effect: (∂ξ_eq/∂T)_p and Δ_r H have the same sign—raising T shifts an exothermic reaction (negative Δ_r H) to the left, opposing the temperature increase.
  • Pressure effect: (∂ξ_eq/∂p)_T and Δ_r V have opposite signs—raising p shifts a reaction with positive Δ_r V (volume increases) to the left, opposing the pressure increase.
  • Le Châtelier's principle: the equilibrium shifts in the direction that tends to oppose the imposed change, but this principle can mislead in certain cases.
  • Common confusion: Le Châtelier's principle fails for adding inert gas at constant volume (little effect on equilibrium despite pressure increase) and for solubility when the differential enthalpy Δ_sol H differs in sign from the integral enthalpy ΔH_m(sol).

🔬 How equilibrium position responds to changes

🔬 Advancement and equilibrium

The advancement ξ of a chemical reaction describes the changes in the amounts of reactants and products from specified initial values.

  • At constant T and p, ξ changes spontaneously in the direction that decreases Gibbs energy G.
  • The system reaches reaction equilibrium at the minimum of G, where ξ = ξ_eq.
  • The value of ξ_eq depends on both T and p; changing either causes the equilibrium to shift to a new position.

📐 Mathematical foundation

The excerpt derives the total differential of Gibbs energy with T, p, and ξ as independent variables:

  • dG = −S dT + V dp + Δ_r G dξ
  • At equilibrium, Δ_r G = 0 and Δ_r S = Δ_r H / T.
  • Substituting these into the differential and setting dΔ_r G = 0 gives:
    • 0 = −(Δ_r H / T) dT + Δ_r V dp + (∂²G/∂ξ²)_{T,p} dξ_eq
  • The second derivative (∂²G/∂ξ²)_{T,p} is positive at the equilibrium minimum.

Why this matters: The positive second derivative ensures that the equilibrium is stable and allows us to solve for how ξ_eq shifts with T or p.

🌡️ Temperature effects on equilibrium

🌡️ The temperature shift rule

Solving the equilibrium differential for constant pressure gives:

  • (∂ξ_eq/∂T)p = (Δ_r H / T) / (∂²G/∂ξ²){T,p}

Key insight: Because the second derivative is positive, (∂ξ_eq/∂T)_p and Δ_r H have the same sign.

Reaction typeΔ_r H signTemperature increaseEquilibrium shift
ExothermicNegativeRaise Tξ_eq decreases (shifts left)
EndothermicPositiveRaise Tξ_eq increases (shifts right)

🔥 Example: exothermic reaction

  • Suppose Δ_r H is negative (exothermic).
  • An increase in temperature causes ξ_eq to decrease: the equilibrium shifts to the left.
  • This is the shift that would reduce the temperature if the reaction were adiabatic—it opposes the imposed change.

Don't confuse: The shift opposes the change, not the condition. Raising T on an exothermic reaction shifts left (toward reactants), even though the system is now hotter.

🗜️ Pressure effects on equilibrium

🗜️ The pressure shift rule

At constant temperature:

  • (∂ξ_eq/∂p)T = −Δ_r V / (∂²G/∂ξ²){T,p}

Key insight: (∂ξ_eq/∂p)_T and Δ_r V have opposite signs.

Reaction volume changeΔ_r V signPressure increaseEquilibrium shift
Volume increasesPositiveRaise pξ_eq decreases (shifts left)
Volume decreasesNegativeRaise pξ_eq increases (shifts right)

📦 Example: volume-increasing reaction

  • If Δ_r V is positive, the volume increases as the reaction proceeds to the right at constant T and p.
  • Increasing the pressure isothermally (by reducing the volume) causes the equilibrium to shift to the left.
  • This is the shift that would reduce the pressure if the reaction occurred at constant T and V—it opposes the imposed change.

⚖️ Le Châtelier's principle and its limits

⚖️ The principle stated

Le Châtelier's principle: When a change is made to a closed system at equilibrium, the equilibrium shifts in the direction that tends to oppose the change.

  • The temperature and pressure shift rules are applications of this principle.
  • The principle is qualitative and intuitive but can be misused.

⚠️ When Le Châtelier's principle misleads

⚠️ Solubility and enthalpy confusion

The excerpt gives the dissolution process B(s) → B(sln):

  • The rate of change of solubility with T, (∂ξ_eq/∂T)_p, has the same sign as the differential enthalpy of solution Δ_sol H at saturation.
  • The sign of Δ_sol H at saturation may differ from the sign of the integral molar enthalpy of solution ΔH_m(sol).
  • Example: sodium acetate at saturation (ξ_sol ≈ 15 mol per kg water):
    • Δ_sol H is positive (endothermic at the margin).
    • ΔH_m(sol) is negative (the overall dissolution of 15 mol is exothermic).
    • Despite the exothermic overall process, solubility increases with temperature because the differential enthalpy is positive.

Don't confuse: The equilibrium shift depends on the differential (marginal) enthalpy at equilibrium, not the integral (total) enthalpy from the initial state.

⚠️ Adding inert gas at constant volume

  • Adding an inert gas at constant V increases the pressure.
  • Le Châtelier's principle might suggest the equilibrium should shift according to Δ_r V.
  • Reality: Adding the inert gas has little effect on the equilibrium position, regardless of Δ_r V.
  • Why: The inert gas affects the activities of reactants and products only slightly (not at all if the gas mixture is ideal), so there is little or no effect on the reaction quotient Q_rxn.
  • Important: The pressure dependence (∂ξ_eq/∂p)_T applies to a closed system; adding inert gas makes the system open with respect to that component.

✅ The rigorous criterion

The rigorous criterion for equilibrium position is always that Q_rxn must equal K, or equivalently, that Δ_r G must be zero.

  • Le Châtelier's principle is a helpful guide but not a substitute for the fundamental equilibrium condition.
  • When in doubt, check whether Q_rxn = K or Δ_r G = 0.
75

Effects of Temperature

12.1 Effects of Temperature

🧭 Overview

🧠 One-sentence thesis

The temperature dependence of equilibrium constants is directly related to the standard molar reaction enthalpy through the van't Hoff equation, allowing noncalorimetric determination of reaction enthalpies.

📌 Key points (3–5)

  • Core relation: The derivative of chemical potential divided by temperature with respect to temperature equals negative partial molar enthalpy divided by temperature squared.
  • Van't Hoff equation: Relates the temperature dependence of the equilibrium constant K to the standard molar reaction enthalpy.
  • Standard state dependence: The exact form of the temperature dependence equation depends on whether solute standard states are based on concentration or mole fraction/molality.
  • Common confusion: The van't Hoff equation (exact) versus the Clausius–Clapeyron equation (approximate)—the former uses standard enthalpies and is exact, while the latter assumes ideal-gas behavior and volume differences.
  • Practical application: Plotting ln K versus 1/T allows determination of standard reaction enthalpy from the slope without calorimetry.

🌡️ Fundamental temperature derivatives

🌡️ Variation of chemical potential ratio with temperature

The excerpt derives how the ratio of chemical potential to temperature changes with temperature:

The partial derivative of μ_i / T with respect to T at constant p and fixed amounts equals negative H_i divided by T squared.

  • Starting from a mathematical operation (product rule), the derivative is: (1/T)(∂μ_i/∂T) - μ_i/T²
  • Using the relation that (∂μ_i/∂T) at constant p and composition equals negative S_i (partial molar entropy)
  • Substituting μ_i = H_i - TS_i (the Gibbs energy relation)
  • Final result: ∂(μ_i/T)/∂T at constant p and composition = -H_i/T²

Don't confuse: This is a general thermodynamic relation, not specific to any particular phase or standard state.

📐 The Gibbs–Helmholtz equation

For a pure substance in a closed system, multiplying the chemical potential relation by amount n gives:

∂(G/T)/∂T at constant p = -H/T²

  • This is called the Gibbs–Helmholtz equation
  • G is the total Gibbs energy, H is the total enthalpy
  • It applies to pure substances, whereas the previous relation applies to species in mixtures

🎯 Standard state temperature dependence

🎯 General form for standard chemical potentials

The excerpt derives how the standard chemical potential ratio varies with temperature by substituting μ_i = μ°_i + RT ln a_i:

  • The derivative d(μ°_i/T)/dT = -H_i/T² - R(∂ ln a_i/∂T) at constant p and composition
  • Because μ°_i/T depends only on T, we can evaluate the right side at any convenient pressure and composition

⚗️ Different standard states require different treatments

Standard state typeActivity term behaviorFinal equation
Gas-phase speciesActivity constant at constant p and compositiond(μ°_i/T)/dT = -H°_i/T²
Mole fraction or molalityActivity constant at constant p and compositiond(μ°_i/T)/dT = -H°_i/T²
Concentration basisActivity changes with T due to volume expansiond(μ°_c,B/T)/dT = -H°_B/T² + Rα_A

Key insight: For gas-phase species or condensed-phase species with standard states based on mole fraction or molality, the activity remains constant when temperature changes at constant pressure and composition, so the partial derivative of ln a_i is zero.

🧪 The concentration basis exception

For solute standard states based on concentration:

  • The concentration c_B = n_B/V changes with temperature even at constant p and composition because volume V changes
  • The derivative (∂ ln(c_B/c°)/∂T) at constant p and composition equals α, the cubic expansion coefficient
  • For an ideal-dilute solution, α is α_A, the cubic expansion coefficient of the pure solvent
  • This adds an extra term Rα_A to the temperature derivative

Example: A solute in aqueous solution with concentration-based standard state will have a temperature dependence that includes the thermal expansion of water.

📊 The van't Hoff equation

📊 Derivation from equilibrium constant

The thermodynamic equilibrium constant K depends only on temperature. Starting from:

  • ln K = -(1/RT) Σ_i ν_i μ°_i (relating K to standard chemical potentials)
  • Taking the derivative: d ln K/dT = -(1/R) Σ_i ν_i d(μ°_i/T)/dT
  • Substituting the standard state temperature derivatives
  • Recognizing that Σ_i ν_i H°_i is the standard molar reaction enthalpy Δ_r H°

📐 Two forms of the van't Hoff equation

General form (including concentration-based standard states):

  • d ln K/dT = Δ_r H°/(RT²) - α_A Σ (ν_i for solutes on concentration basis)
  • The sum includes only solute species whose standard states are based on concentration

Simplified form (no concentration-based standard states):

  • d ln K/dT = Δ_r H°/(RT²)
  • Equivalently: Δ_r H° = RT² d ln K/dT
  • Or using the identity d(1/T) = -(1/T²)dT: Δ_r H° = R d ln K/d(1/T)

🔬 Practical application for determining reaction enthalpy

The van't Hoff equation allows evaluation of the standard molar reaction enthalpy by a noncalorimetric method from the temperature dependence of ln K.

Procedure:

  • Plot ln K versus 1/T
  • According to the equation Δ_r H° = R d ln K/d(1/T), the slope of the curve at any value of 1/T equals -Δ_r H°/R at the corresponding temperature T
  • This provides a way to measure reaction enthalpy without using a calorimeter

Example: If you measure equilibrium constants at several temperatures, plotting ln K against 1/T gives a curve whose slope directly reveals the standard reaction enthalpy.

🔄 Comparison with Clausius–Clapeyron equation

🔄 Similarities and differences for vaporization

The excerpt provides a detailed comparison table for liquid vaporization:

FeatureClausius–Clapeyronvan't Hoff
FormΔ_vap H = R d ln(p/p°)/d(1/T)Δ_vap H° = R d ln K/d(1/T)
ExactnessApproximate (assumes V_m(g) >> V_m(l) and ideal-gas behavior)Exact relation
Enthalpy meaningDifference of molar enthalpies of real gas and liquid at saturation vapor pressureDifference of molar enthalpies of ideal gas and liquid at standard pressure p°
Pressure meaningp is the saturation vapor pressure of the liquidK equals a(g)/a(l) = (f/p°)/γ(l), only approximately equal to p/p°

Don't confuse:

  • The Clausius–Clapeyron equation is derived with approximations and applies to real substances at their saturation conditions
  • The van't Hoff equation is exact and uses standard states (ideal gas for vapor, specified standard state for liquid)
  • Both have similar mathematical forms but represent different physical quantities
76

Solvent Chemical Potentials from Phase Equilibria

12.2 Solvent Chemical Potentials from Phase Equilibria

🧭 Overview

🧠 One-sentence thesis

Physical measurements of phase equilibria—such as freezing-point depression, osmotic pressure, or vapor pressure—allow us to determine the difference in solvent chemical potential between pure solvent and solution, which in turn enables calculation of solute activity coefficients.

📌 Key points

  • Core goal: measure the chemical potential difference (μ*_A − μ_A) between pure solvent and solvent in solution at constant T and p; this difference is always positive because solute lowers the solvent's chemical potential.
  • Why it matters: once we know (μ*_A − μ_A), we can calculate the osmotic coefficient φ_m, and from that integrate to find the solute activity coefficient.
  • Three experimental methods: freezing-point depression, osmotic pressure, and isopiestic vapor-pressure measurements all provide routes to (μ*_A − μ_A).
  • Common confusion: freezing-point depression and osmotic pressure are not direct measurements of activity coefficients; they measure the solvent chemical potential difference, which is then converted via the osmotic coefficient.
  • Typical applications: freezing-point and vapor-pressure methods are often used for electrolyte solutions; osmotic pressure is especially useful for macromolecule solutions.

🔗 The measurement chain

🎯 What we measure and what we want

The excerpt describes a multi-step procedure:

  1. Measure (μ*_A − μ_A) over a range of molality at constant T and p using phase-equilibrium experiments.
  2. Convert these values to the osmotic coefficient φ_m using the relation:

    φ_m = (μ*_A − μ_A) / (ν R T M_A m_B) where ν = 1 for a nonelectrolyte and equals the number of ions per formula unit for an electrolyte; M_A is the molar mass of the solvent; m_B is the molality of the solute.

  3. Integrate from infinite dilution to the molality of interest to evaluate the solute activity coefficient.

🧩 Why the solvent chemical potential difference is key

  • The presence of solute lowers the solvent's chemical potential: μ_A (in solution) < μ*_A (pure solvent).
  • Therefore (μ*_A − μ_A) is always positive.
  • This difference encodes the thermodynamic effect of the solute on the solvent.
  • Example: in a dilute aqueous solution, the chemical potential of water in the solution is lower than that of pure water at the same T and p.

🧊 Freezing-point depression method

🧊 Basic principle

  • When a binary solution (solvent A + solute B) is cooled at constant pressure and composition, the solid that first appears is assumed to be pure A (e.g., ice from an aqueous solution).
  • The freezing point T_f of the solution is lower than the freezing point T*_f of the pure solvent.
  • This depression is a consequence of the lowering of μ_A by the solute.

📐 How the measurement works

The excerpt outlines the logic (details continue in a figure not fully shown):

  • Both T_f (solution freezing point) and T*_f (pure solvent freezing point) are measured experimentally.
  • Let T_0 be a temperature of interest equal to or greater than T*_f.
  • The goal is to determine (μ_A(l, T_0) − μ_A(sln, T_0)), where μ_A(l, T_0) is the chemical potential of pure liquid solvent and μ_A(sln, T_0) is the chemical potential of solvent in the solution.
  • A figure (Figure 12.1, not fully reproduced here) plots μ_A/T for the solvent in the pure solid phase, pure liquid phase, and fixed-composition solution as functions of T at constant p.

🔬 Additional data needed

  • The excerpt states that freezing-point measurements must be combined with additional data from calorimetric measurements to evaluate (μ*_A − μ_A).
  • This implies that knowing T_f alone is not sufficient; enthalpy or heat-capacity data are also required.

🧪 Typical use case

  • Freezing-point depression is often used for electrolyte solutions.
  • Example: measuring the freezing point of a salt solution to determine the osmotic coefficient and then the mean ionic activity coefficient.

💧 Osmotic pressure method

💧 What it measures

  • Osmotic pressure is another experimental route to determine (μ*_A − μ_A).
  • The excerpt does not detail the procedure but states that osmotic pressure is especially useful for solutions of macromolecules.

🧬 Why macromolecules?

  • Macromolecules (e.g., proteins, polymers) are large solutes with low molality but significant effects on solvent chemical potential.
  • Osmotic pressure measurements are sensitive enough to detect these effects even at low concentrations.

🌡️ Comparison with van't Hoff and Clausius–Clapeyron equations

📊 Van't Hoff equation for equilibrium constants

The excerpt also discusses the temperature dependence of the equilibrium constant K, which is related but distinct from the solvent chemical potential methods:

  • The van't Hoff equation relates the temperature derivative of ln K to the standard molar reaction enthalpy:

    Δ_r H° = R d(ln K) / d(1/T) (when no solute standard states are based on concentration)

  • This is an exact relation and allows noncalorimetric determination of Δ_r H° from the temperature dependence of K.
  • Example: plot ln K versus 1/T; the slope at any point equals −Δ_r H° / R at the corresponding temperature.

📊 Clausius–Clapeyron vs van't Hoff for vaporization

The excerpt includes a table comparing the two equations for vaporization of a liquid:

FeatureClausius–Clapeyronvan't Hoff
EquationΔ_vap H ≈ R d(ln(p/p°)) / d(1/T)Δ_vap H° = R d(ln K) / d(1/T)
AssumptionsAssumes V_m(gas) ≫ V_m(liquid) and ideal-gas behaviorExact relation
Enthalpy meaningDifference of molar enthalpies of real gas and liquid at saturation vapor pressureDifference of molar enthalpies of ideal gas and liquid at standard pressure p°
Pressure meaningp is the saturation vapor pressure of the liquidK = a(gas)/a(liquid) = (f/p°)/γ(liquid), only approximately equal to p/p°

🔍 Don't confuse

  • The Clausius–Clapeyron equation is an approximation for real vaporization processes.
  • The van't Hoff equation is exact for the standard reaction enthalpy when applied to equilibrium constants.
  • Both involve temperature derivatives, but they refer to different thermodynamic quantities (real vs standard states).

🔄 Isopiestic vapor-pressure method

🔄 Brief mention

  • The excerpt states that the isopiestic vapor-pressure method was described in an earlier section (Sec. 9.6.4, not included here).
  • This method is also used to evaluate (μ*_A − μ_A) and is often used for electrolyte solutions.
  • The term "isopiestic" refers to equal vapor pressure: solutions are equilibrated until they have the same vapor pressure, allowing comparison of solvent chemical potentials.
77

12.3 Binary Mixture in Equilibrium with a Pure Phase

12.3 Binary Mixture in Equilibrium with a Pure Phase

🧭 Overview

🧠 One-sentence thesis

Measuring phase equilibria such as freezing points and osmotic pressures allows us to determine the difference in solvent chemical potential between pure solvent and solution, which in turn enables calculation of solute activity coefficients.

📌 Key points (3–5)

  • Goal of the methods: determine the chemical potential difference (μ*_A − μ_A) between pure solvent and solvent in solution at a given temperature and pressure.
  • Why this matters: the chemical potential difference is used to calculate the osmotic coefficient φ_m, which then allows evaluation of solute activity coefficients.
  • Two main experimental methods: freezing-point depression and osmotic pressure measurements provide different routes to the same chemical potential difference.
  • Common confusion: the freezing point T_f of the solution is lower than the pure solvent freezing point T*_f because the solute lowers the solvent's chemical potential, not because of direct temperature effects.
  • Integration path logic: freezing-point methods require integrating enthalpy changes along a thermodynamic path connecting pure phases and solution phases at different temperatures.

🔬 The measurement strategy

🎯 What needs to be determined

The chemical potential difference μ*_A − μ_A: the difference between the chemical potentials of the pure solvent and the solvent in the solution at the temperature and pressure of interest.

  • This difference is always positive because the presence of solute reduces the solvent's chemical potential.
  • Once this difference is known over a range of molality at constant T and p, it can be converted to the osmotic coefficient φ_m.

🔄 The calculation sequence

The excerpt describes a three-step process:

  1. Determine μ*_A − μ_A over a range of molality at constant T and p
  2. Convert these values to φ_m using the equation: φ_m = (μ*_A − μ_A) / (ν RT M_A m_B), where ν = 1 for nonelectrolytes and equals the number of ions per formula unit for electrolytes
  3. Evaluate the solute activity coefficient by integration from infinite dilution to the molality of interest

🧪 Which methods are used for which systems

MethodTypical application
Freezing-point depressionOften used for electrolyte solutions
Isopiestic vapor-pressureOften used for electrolyte solutions
Osmotic pressureEspecially useful for solutions of macromolecules

❄️ Freezing-point depression method

🧊 The basic setup

  • Consider a binary solution of solvent A and solute B.
  • When cooled at constant pressure and composition, the solid that first appears is pure A (e.g., ice from dilute aqueous solution).
  • The temperature at which solid A first appears is T_f, the freezing point of the solution.
  • This temperature is lower than the pure solvent freezing point T*_f.

📉 Why the freezing point is lowered

  • The lowering of T_f below T*_f is a consequence of the lowering of μ_A by the presence of the solute.
  • At the freezing point T_f, the chemical potential of A in the solution equals the chemical potential of pure solid A.
  • At the pure solvent freezing point T*_f, the chemical potential of pure liquid A equals the chemical potential of pure solid A.
  • Don't confuse: the lower freezing point is not a direct effect of the solute on temperature, but rather reflects the equilibrium condition when the solvent's chemical potential has been reduced.

🛤️ The integration path

The excerpt describes an integration path (labeled abcde in Figure 12.1) that connects different states:

  • The goal is to find the difference in μ_A/T at temperature T_0 between pure liquid (point e) and solution (point a).
  • The path integrates the slope d(μ_A/T)/dT over temperature.
  • The slope at any point equals −H_A/T², where H_A is the partial molar enthalpy in that phase.

The integration proceeds through:

  • Solution phase from T_0 down to T*_f
  • Solution phase from T*_f down to T_f
  • Pure solid phase from T_f up to T*_f
  • Pure liquid phase from T*_f up to T_0

🔥 Required enthalpy data

Differential enthalpy of solution (Δ_sol,A H)

The molar enthalpy difference H_A(sln) − H*_A(s): the enthalpy change when solid A dissolves in the solution at constant T and p.

  • This can be measured calorimetrically at the equilibrium temperature T_f.
  • The excerpt assumes it varies linearly with temperature: Δ_sol,A H(T) = Δ_sol,A H(T_f) + Δ_sol,A C_p (T − T_f).
  • The differential heat capacity of solution Δ_sol,A C_p = C_p,A(sln) − C_p,A(s) is treated as constant.

Differential enthalpy of dilution (Δ_dil H)

  • This is the quantity H_A(sln) − H*_A(l), the enthalpy change when pure liquid A is diluted in the solution.
  • Can be measured calorimetrically at any temperature higher than T*_f.

📐 The final equation

The excerpt provides the complete formula for the chemical potential difference at temperature T_0:

μ*_A(l, T_0) − μ_A(sln, T_0) equals:

  • T_0 times Δ_sol,A H(T_f) divided by T_f
  • minus T_0 times Δ_sol,A C_p times (1/T_f − 1/T*_f)
  • plus T_0 times Δ_sol,A C_p times ln(T*_f / T_f)
  • plus T_0 times the integral from T*_f to T_0 of (Δ_dil H / T²) dT

This equation combines the measured freezing points (T_f and T*_f) with calorimetric data to yield the desired chemical potential difference.

🧫 Osmotic pressure method

🔬 The apparatus

The excerpt mentions an apparatus (Figure 12.2) but the description is incomplete in the provided text. It indicates:

  • A membrane permeable only to solvent A (shown as a dashed line).
  • Pure solvent A(l) at pressure p' on one side.
  • Solution (A + B)(sln) at pressure p'' on the other side.

The osmotic pressure method provides an alternative experimental route to determine the same chemical potential difference, particularly useful for macromolecule solutions.

78

Solvent Chemical Potentials from Phase Equilibria

12.4 Colligative Properties of a Dilute Solution

🧭 Overview

🧠 One-sentence thesis

Measuring phase equilibria—such as freezing-point depression and osmotic pressure—provides practical methods to determine the difference in solvent chemical potential between a pure solvent and a solution.

📌 Key points (3–5)

  • Two experimental methods: freezing-point measurements and osmotic-pressure measurements both yield the difference in solvent chemical potential between pure solvent and solution.
  • Freezing-point approach: integrates enthalpy differences along a path from the solution freezing point to the pure-solvent freezing point to find the chemical potential difference.
  • Osmotic pressure approach: uses a semipermeable membrane to establish equilibrium between pure solvent and solution at different pressures; the pressure difference (osmotic pressure) relates directly to the chemical potential difference.
  • Common confusion: osmotic pressure is an intensive property of the solution at a given temperature, pressure, and composition—it exists whether or not the additional pressure is actually applied, just as a solution has a freezing point even when not at that temperature.
  • Why it matters: these methods allow experimental determination of solvent activity and activity coefficients, which are central to understanding solution behavior.

🧊 Freezing-point method

🧊 Core idea

  • The solvent chemical potential is equal in coexisting phases at equilibrium (solution/solid at the solution freezing point T_f, pure liquid/solid at the pure-solvent freezing point T*_f).
  • By integrating the slope of chemical potential curves (which depend on enthalpy) along a path from one freezing point to the other, we can calculate the chemical potential difference at any temperature T_0.

🔢 The integration path

  • The excerpt describes an integration path abcde at constant pressure (Figure 12.1).
  • The path connects:
    • Point a: solution at T_0
    • Point e: pure liquid at T_0
  • The integration passes through the two freezing points (T_f for the solution, T*_f for the pure solvent) and involves solid and liquid phases.
  • The slope of each curve is given by the enthalpy divided by temperature squared (from an earlier equation).

📐 The working equation

The final result after combining integrals and substituting enthalpy differences is:

Chemical potential difference:
μ_A(pure liquid, T_0) − μ_A(solution, T_0) = T_0 · [ΔH_sol,A(T_f) / T_f − ΔC_p,sol,A · (1/T_f − 1/T_f)] + T_0 · ΔC_p,sol,A · ln(T_f / T_f) + T_0 · ∫[T*_f to T_0] (ΔH_dil / T²) dT

Where:

  • ΔH_sol,A: molar differential enthalpy of solution of solid A in the solution (H_A(solution) − H*_A(solid)).
  • ΔC_p,sol,A: molar differential heat capacity of solution (C_p,A(solution) − C_p,A(solid)), treated as constant.
  • ΔH_dil: molar differential enthalpy of dilution of the solvent in the solution (H_A(solution) − H*_A(liquid)).

🔬 Practical assumptions

  • Linear temperature dependence: ΔH_sol,A is assumed to be a linear function of temperature, with slope ΔC_p,sol,A.
  • Calorimetric measurements: ΔH_sol,A is measured at the equilibrium freezing point T_f; ΔH_dil can be measured at any temperature above T*_f.
  • This approach requires knowledge of enthalpy differences over a temperature range, which is obtained from calorimetry.

💧 Osmotic-pressure method

💧 What osmotic pressure is

Osmotic pressure (Π): the additional pressure that must be exerted on a solution (compared to the pressure p' of pure solvent at the same temperature) to establish equilibrium with no net solvent flow through a semipermeable membrane.

  • The membrane is permeable to solvent but impermeable to solute.
  • At equilibrium: p'' = p' + Π, where p'' is the solution pressure and p' is the pure-solvent pressure.

🧪 The apparatus and equilibrium condition

  • Setup (Figure 12.2): two liquid phases separated by a semipermeable membrane, with movable pistons to control pressure.
    • Phase α: pure solvent at pressure p'.
    • Phase β: solution at pressure p''.
  • Equilibrium conditions:
    • Temperature is the same in both phases.
    • Solvent chemical potential is the same in both phases: μ_A(p'') = μ_A(p' + Π) = μ*_A(p').
    • Pressures can be unequal because the membrane constrains solute movement.

🌊 Osmosis phenomenon

  • Starting condition: both phases at the same temperature and pressure.
  • What happens: μ_A is lower in the solution than in the pure liquid, so solvent spontaneously flows through the membrane from pure solvent to solution.
  • Osmosis: this spontaneous flow of solvent through the membrane (Greek for "push").
  • Stopping the flow: increase the pressure on the solution side until μ_A in the solution equals μ*_A in the pure solvent; the required pressure increase is the osmotic pressure Π.

📊 Key properties of osmotic pressure

PropertyDescription
IntensiveDepends on temperature, pressure, and composition of the solution
Reference stateDefined relative to pure solvent at pressure p' and temperature T
Always existsA solution has an osmotic pressure even when the additional pressure is not actually applied, just as a solution has a freezing point even when not at that temperature
Practical toleranceThe membrane need not be perfectly impermeable to solute; it only needs to establish solvent equilibrium on a short time scale with negligible solute transfer

🔗 Relation to chemical potential

The equilibrium condition gives:

μ_A(p' + Π) = μ*_A(p')

This relation allows derivation of an expression for μ*_A(p') − μ_A(p') as a function of the osmotic pressure Π, using the dependence of chemical potential on pressure.

🔄 Connection to solvent activity

🔄 Activity coefficient from chemical potential

  • A measurement of Δμ_A = μ*_A − μ_A also gives the solvent activity coefficient γ_A, based on the pure-solvent reference state.
  • The relation is: μ_A = μ*_A + RT ln(γ_A · x_A), where x_A is the mole fraction of solvent.
  • Don't confuse: this is the activity coefficient for the solvent, not the solute; it measures how the solvent's chemical potential in solution deviates from ideal-solution behavior.

🎯 Why these methods matter

  • Both freezing-point and osmotic-pressure measurements provide experimental access to solvent chemical potentials.
  • These quantities are fundamental to understanding solution thermodynamics, including deviations from ideality and the behavior of real solutions.
  • The methods complement each other: freezing-point measurements involve temperature changes and enthalpy integration; osmotic-pressure measurements involve pressure changes at constant temperature.
79

Solid–Liquid Equilibria

12.5 Solid–Liquid Equilibria

🧭 Overview

🧠 One-sentence thesis

Solid–liquid equilibria describe how freezing points and solubilities vary with composition, and ideal behavior can be predicted when the liquid mixture behaves ideally and the enthalpy of solution equals the enthalpy of fusion.

📌 Key points (3–5)

  • Freezing-point and solubility curves are equivalent: the composition of a solution at its freezing point is the same as the solubility of the solid in that solution.
  • Ideal solubility depends only on the solid's properties: for an ideal liquid mixture, solubility depends on the solid's melting point and enthalpy of fusion, not on the solvent identity.
  • Common confusion—freezing vs. solubility: these describe the same equilibrium from different perspectives (temperature as a function of composition vs. composition as a function of temperature).
  • Solid compounds form freezing-point maxima: when a solid compound forms from two components, the freezing-point curve has a maximum at the compound's composition.
  • Electrolyte solubility involves the solubility product: the equilibrium constant relates to ion molalities and mean ionic activity coefficients in saturated solutions.

🧊 Freezing behavior of ideal mixtures

🧊 Freezing-point curves for ideal liquid mixtures

A freezing-point curve shows freezing point as a function of liquid composition; a solubility curve shows composition of a solution in equilibrium with a pure solid as a function of temperature.

  • These are two ways of describing the same physical situation.
  • Example: the composition of an aqueous solution at the freezing point is the mole fraction solubility of ice in the solution.

📐 Equation for ideal freezing-point depression

The general relation for an ideal binary liquid mixture in equilibrium with pure solid A is:

  • The derivative of freezing temperature with respect to mole fraction equals: (R times T-freezing squared) divided by (mole fraction times enthalpy of solution).
  • When the enthalpy of solution is constant and equal to the enthalpy of fusion, integration gives: ln(x_A) = (enthalpy of fusion / R) × (1/T-melting-pure − 1/T-freezing).

📊 Shape of freezing-point curves

The shape depends on the ratio of enthalpy of fusion to RT at the melting point:

Ratio valueCurve shapeBehavior
Greater than 2Concave downwardTypical behavior
Less than 2Concave upward at low x_BUnusual behavior
  • Example: Benzene–toluene mixtures show good agreement with ideal behavior over a wide composition range; benzene–cyclohexane mixtures agree only in the dilute region.

🧪 Solubility of solid nonelectrolytes

🧪 Defining solubility

A solution is saturated with respect to a solid when it can exist in transfer equilibrium with pure solid at the same temperature and pressure.

Solubility is the maximum value of solute mole fraction, concentration, or molality that can exist without spontaneous precipitation.

  • At standard pressure, solubility (x_B) equals the equilibrium constant divided by the activity coefficient.
  • If the activity coefficient is unity (dilute solution), solubility equals the equilibrium constant directly.

🌡️ Temperature dependence of solubility

The standard molar enthalpy of solution can be found from how solubility changes with temperature:

  • Enthalpy of solution = R × T² × (derivative of ln(solubility) with respect to T).
  • If solubility increases with temperature → solution process is endothermic (positive enthalpy).
  • If solubility decreases with temperature → solution process is exothermic (negative enthalpy).
  • Don't confuse: this applies to low-solubility solids; the general relation is more complex.

🎯 Ideal solubility

Ideal solubility is calculated assuming (1) the liquid is an ideal mixture and (2) the enthalpy of solution equals the enthalpy of fusion.

The ideal solubility equation is: ln(x_B) = (enthalpy of fusion / R) × (1/T-melting − 1/T).

Key properties of ideal solubility:

  • Independent of solvent type: depends only on the solid's properties.
  • Increases with temperature: always becomes more soluble at higher T.
  • Lower for high-melting solids: at a given temperature, solids with higher melting points have lower ideal solubility (for similar enthalpies of fusion).

Example: Solid benzene has close to ideal solubility in liquid toluene at temperatures not lower than about 20 K below benzene's melting point.

🔗 Solid compounds from mixture components

🔗 What are solid compounds?

A solid compound (or stoichiometric addition compound) is a solid containing both components in a fixed proportion.

  • Examples: salt hydrates (salts with fixed waters of hydration) and certain metal alloys.
  • The liquid mixture has variable composition; the solid compound has fixed composition.
  • Assumption: in the liquid phase, the compound is completely dissociated into components.

📈 Freezing-point maximum

When a solid compound forms, the freezing-point curve has a maximum at the composition matching the solid compound.

  • At the maximum, the liquid and solid have the same composition.
  • The slope of the freezing-point curve is zero at this composition.
  • The slope decreases as the mole fraction increases toward the compound composition.

Example: A molten metal mixture of Zn and Mg solidifies to the alloy Zn₂Mg. The freezing-point curve has a maximum at x_Zn = 2/3 (the compound composition) at 861 K.

🧮 Equation for solid compound equilibria

For an ideal liquid mixture in equilibrium with a solid compound A_a B_b, assuming constant enthalpy equal to the enthalpy of fusion:

  • 1/T-freezing-initial = 1/T-melting-compound + (R / enthalpy of fusion) × [a × ln(x_A-final / x_A-initial) + b × ln(x_B-final / x_B-initial)].
  • This equation closely approximates the freezing behavior of some metal alloy systems.

⚡ Solubility of solid electrolytes

⚡ Solubility product

For an ionic solid in equilibrium with its dissolved ions: M_νC X_νA (s) ⇌ νC M^(zC) (aq) + νA X^(zA) (aq)

The solubility product K_s is the thermodynamic equilibrium constant for this dissolution equilibrium.

The general expression is:

  • K_s = (pressure factor) × (mean ionic activity coefficient)^ν × (m_cation / m°)^νC × (m_anion / m°)^νA
  • Where ν = νC + νA is the total number of ions per formula unit.

🧂 No common ion case

When the salt dissolves in pure water (no common ions):

  • K_s = (pressure factor) × (mean ionic activity coefficient)^ν × (νC^νC × νA^νA) × (m_B / m°)^ν
  • Where m_B is the solubility expressed as molality.
  • At low ionic strength, the mean ionic activity coefficient can be estimated from the Debye–Hückel limiting law.

🔄 Common ion effect

The common ion effect: adding a salt with a common ion decreases the solubility of a sparingly-soluble salt.

  • Since K_s is constant at fixed T and p, increasing the concentration of a common ion forces the solubility to decrease.
  • Example: Adding a soluble salt M'_νC' Y_νA' (sharing the cation M^zC) decreases the solubility of the sparingly-soluble salt M_νC X_νA.

🌊 Salting-in effect

Salting-in effect: the solubility of a sparingly-soluble salt increases when a second salt lacking a common ion is dissolved.

  • This occurs because the added salt decreases the mean ionic activity coefficient.
  • Since K_s is constant, a decrease in activity coefficient requires an increase in solubility.

🌡️ Temperature dependence

The temperature dependence of the solubility product is:

  • d(ln K_s) / dT = (standard enthalpy of solution) / (R × T²)
  • At standard pressure, the standard enthalpy of solution equals the molar enthalpy of solution at infinite dilution.
80

Liquid–Liquid Equilibria

12.6 Liquid–Liquid Equilibria

🧭 Overview

🧠 One-sentence thesis

When two liquids cannot mix in all proportions, they form two coexisting liquid phases with compositions determined solely by the substances' identities, temperature, and pressure, and the distribution of a third solute between these phases follows predictable equilibrium relationships.

📌 Key points (3–5)

  • Partial miscibility: two liquids that cannot mix in all proportions form two coexisting liquid phases with fixed compositions at equilibrium.
  • Solubility definition: the maximum amount of one liquid that can dissolve in another before phase separation occurs.
  • Two ways to treat solubility: using either a solute standard state (when solubility is low) or a pure-liquid reference state (when treating as a mixture constituent).
  • Common confusion: solubility can show a minimum with temperature—below the minimum temperature the enthalpy of solution is negative, above it is positive.
  • Nernst distribution law: a third solute distributes between two partially-miscible solvents with a constant mole fraction ratio at dilute concentrations.

🧪 Partial miscibility and phase separation

🧪 What partial miscibility means

Partially miscible: two pure liquids that are unable to mix in all proportions.

  • When two partially miscible liquids contact each other and reach equilibrium, the result is two coexisting liquid mixtures with different compositions.
  • Liquids are never actually completely immiscible—even mercury and water dissolve trace amounts of each other, though the amounts may be too small to measure.
  • Example: water (polar) and benzene (nonpolar) form two phases—an aqueous phase with dissolved benzene and a benzene phase with dissolved water.

🔒 Fixed compositions at equilibrium

  • At a given temperature and pressure, the mole fraction compositions of both phases are fixed.
  • The compositions depend only on the identity of the substances, temperature, and pressure.
  • This follows from the Gibbs phase rule: a two-component, two-phase system at equilibrium has only two independent intensive variables.

🔥 Why phase separation occurs

  • Phase separation is usually the result of positive deviations from Raoult's law.
  • Typically occurs when one substance is polar and the other is nonpolar.
  • Phase separation often happens when the temperature is lowered.

💧 Solubility of one liquid in another

💧 Defining solubility

Solubility of B in A: the maximum amount of B that can dissolve without phase separation in a given amount of A at the given temperature and pressure.

  • Can be expressed as the mole fraction of B in the phase at the point of phase separation.
  • Adding any more B beyond this point results in two coexisting liquid phases of fixed composition.
  • Experimentally determined from the cloud point—the point during titration where persistent turbidity is observed.

🎯 Two approaches to calculating solubility

ApproachWhen to useKey assumptionSolubility formula
Solute standard stateLow mutual solubilitiesB has low mole fraction in phase ' and close to 1 in phase "x'_B ≈ K (equilibrium constant)
Pure-liquid reference stateTreating as mixture constituentB-rich phase is almost pure liquid Bx'_B ≈ 1/γ'_B
  • In both approaches, pressure factors and activity coefficients are close to 1 for low solubilities.
  • The solute standard state treats B as a solute in the dilute phase and as a mixture constituent in the concentrated phase.
  • The pure-liquid reference state uses the same reference for B in both phases.

🌡️ Temperature dependence of solubility

The temperature dependence is given by:

  • d(ln x'_B)/dT = ΔH°_sol,B / (R T²)
  • Where ΔH°_sol,B is the molar enthalpy change for transferring pure liquid solute to the solution at infinite dilution.

Don't confuse: solubility doesn't always increase with temperature.

  • Example: n-butylbenzene in water shows a minimum solubility at about 12°C.
  • Below 12°C, ΔH°_sol,B is negative (dissolution releases heat).
  • Above 12°C, ΔH°_sol,B is positive (dissolution absorbs heat).

🌊 Environmental chemistry application

For nonpolar liquid solutes with very low water solubility:

  • Use a pure-liquid reference state.
  • The aqueous solution is essentially at infinite dilution.
  • The activity coefficient γ'_B is a limiting activity coefficient (activity coefficient at infinite dilution).
  • γ'_B is much greater than 1.
  • Solubility: x'_B ≈ 1/γ'_B.
  • Can also relate solubility to Henry's law constant: x'_B ≈ p_B / k'_H,B, where p_B is the vapor pressure of pure solute.

⚖️ Distribution of a solute between two solvents

⚖️ Setting up the distribution problem

  • Start with a two-component system of two equilibrated liquid phases ' and ".
  • Add a small quantity of a third component C.
  • C distributes itself between the two phases.
  • Treat C as a solute in both phases.

📐 The Nernst distribution law

Define K' as the ratio of mole fractions of C in the two phases at equilibrium:

  • K' = x'_C / x"_C

Nernst distribution law: at a fixed temperature and pressure, if x_C is low enough in both phases for the activity coefficients to be close to unity, K' becomes a constant.

  • The equilibrium constant and pressure factors are constants at fixed T and p.
  • When activity coefficients γ'_x,C and γ"_x,C are close to 1, the mole fraction ratio is constant.
  • Since molality and concentration are proportional to mole fraction in dilute solutions, the ratios m'_C/m"_C and c'_C/c"_C also approach constant values.

🔄 Partition coefficient

Partition coefficient (or distribution coefficient): the ratio of concentrations c'_C / c"_C in the two phases.

  • Applies in the dilute solution limit.
  • At infinite dilution of C, the two phases have the compositions that exist when only the original two components A and B are present.

🔀 Beyond dilute solutions

  • As x'_C and x"_C increase beyond the dilute region, the ratios x'_B/x'_A and x"_B/x"_A may change.
  • Continued addition of C may increase the mutual solubilities of A and B.
  • When enough C has been added, a single liquid phase containing all three components may form.
  • Don't confuse: the distribution coefficient is only constant in the dilute regime; adding more solute can eventually change the phase behavior entirely.
81

Membrane Equilibria

12.7 Membrane Equilibria

🧭 Overview

🧠 One-sentence thesis

When two liquid phases are separated by a semipermeable membrane, equilibrium requires equal temperature and equal chemical potential for permeable species, but the phases typically have different pressures and (if charged species are involved) different electric potentials.

📌 Key points (3–5)

  • Core equilibrium condition: permeable species must have the same chemical potential in both phases; the two phases need not have the same pressure.
  • Osmotic membrane equilibrium: when a membrane is permeable only to solvent, the pressure difference between phases is related to their osmotic pressures.
  • Donnan membrane equilibrium: when charged species are involved and the membrane is impermeable to certain ions (e.g., a polyelectrolyte), equilibrium produces both a pressure difference and an electric potential difference (Donnan potential) across the membrane.
  • Common confusion: in osmotic equilibrium, only pressure differs; in Donnan equilibrium, both pressure and electric potential differ, and ion concentrations are unequal even though the salt's chemical potential is equal.
  • Why it matters: these principles explain osmotic pressure measurements, equilibrium dialysis for studying ligand binding, and membrane potentials in biological systems.

🧪 General membrane equilibrium principles

🧪 What a semipermeable membrane does

  • A semipermeable membrane separates two liquid phases and is permeable to certain species but impermeable to others.
  • In principle, the membrane allows some species to pass freely while blocking others.
  • The excerpt assumes that during observation, permeable species quickly reach transfer equilibrium and only negligible amounts of impermeable species cross the membrane.

⚖️ Equilibrium conditions

In an equilibrium state, both phases must have the same temperature, and any species to which the membrane is permeable must have the same chemical potential in both phases.

  • Key point: the two phases need not and usually do not have the same pressure.
  • This is different from ordinary phase equilibrium without a membrane, where pressure is also equal.
  • Example: if solvent A can pass through the membrane, then the chemical potential of A in phase ' equals the chemical potential of A in phase ".

🌊 Osmotic membrane equilibrium

🌊 What osmotic membrane equilibrium is

An osmotic membrane equilibrium: an equilibrium state in a system with two solutions of the same solvent and different solute compositions, separated by a membrane permeable only to the solvent.

  • This is the situation in an apparatus that measures osmotic pressure (referenced as Fig. 12.2 on page 373).
  • Only the solvent can cross the membrane; solutes cannot.

🔢 Pressure difference in osmotic equilibrium

  • The solvent has equal chemical potential in both phases: μ"_A(p") = μ'_A(p').
  • The dependence of chemical potential on pressure is given by (∂μ_A/∂p) at constant T and composition = V̄_A, where V̄_A is the partial molar volume of A.
  • Integrating this relation from the pressure of phase ' to that of phase " gives:
    • μ"_A(p") = μ"_A(p') + V̄"_A(p" − p')
  • Equating the two expressions for μ"_A(p") and rearranging yields the pressure difference:
    • p" − p' = [μ'_A(p') − μ"_A(p')] / V̄"_A
  • This pressure difference can be related to the osmotic pressures of the two phases.
  • Using the relation μ_A(p) = μ*_A(p) − V̄_A Π(p), the pressure difference becomes:
    • p" − p' = Π"(p') − (V̄'_A / V̄"_A) Π'(p')
  • Example: if phase " has higher osmotic pressure, it must also have higher actual pressure to maintain equilibrium.

🧬 Equilibrium dialysis

🧬 What equilibrium dialysis measures

Equilibrium dialysis: a useful technique for studying the binding of a small uncharged solute species (a ligand) to a macromolecule.

  • The macromolecule solution is placed on one side of a membrane through which it cannot pass.
  • A solution without the macromolecule is on the other side.
  • The ligand is allowed to come to transfer equilibrium across the membrane.

🔬 How it works

  • If the same solute standard state is used for the ligand in both solutions, at equilibrium the unbound ligand must have the same activity in both solutions.
  • Measurements include:
    • Total ligand molality in the macromolecule solution
    • Ligand molality in the other solution
  • Combined with estimated values of the unbound ligand activity coefficients, these measurements allow the amount of ligand bound per macromolecule to be calculated.
  • Don't confuse: the ligand activity is equal in both phases, but the total ligand concentration differs because some ligand is bound to the macromolecule.

⚡ Donnan membrane equilibrium

⚡ What makes Donnan equilibrium different

Donnan membrane equilibrium: the equilibrium state when one solution contains certain charged solute species that are unable to pass through the membrane, whereas other ions can pass through.

  • This is more complicated than osmotic membrane equilibrium.
  • Usually if the membrane is impermeable to one kind of ion, an ion species to which it is permeable achieves transfer equilibrium only when the phases have different pressures and different electric potentials.

Donnan potential: the resulting electric potential difference across the membrane.

  • This phenomenon is related to membrane potentials important in nerve and muscle cells (though living cells are not in equilibrium states).
  • A Donnan potential can be measured electrically using silver-silver chloride electrodes connected to both phases through salt bridges (with some uncertainty due to unknown liquid junction potentials).

🧪 General setup and conditions

Consider two solution phases ' and " separated by a semipermeable membrane:

  • Both phases contain a dissolved salt (solute B) with ν₊ cations and ν₋ anions per formula unit.
  • The membrane is permeable to these ions.
  • Phase " also contains a protein or other polyelectrolyte with a net positive or negative charge, together with counterions of the opposite charge.
  • The counterions are the same species as the cation or anion of the salt.
  • The membrane is impermeable to the polyelectrolyte (perhaps because membrane pores are too small).

Transfer equilibrium condition for salt B: μ'_B = μ"_B

This leads to the relation:

  • (γ'_m,B)^(ν₊+ν₋) (m'₊)^ν₊ (m'₋)^ν₋ = (γ"_m,B)^(ν₊+ν₋) (m"₊)^ν₊ (m"₋)^ν₋

where γ_m,B is the mean ionic activity coefficient and m₊, m₋ are the molalities of cations and anions.

🔋 Donnan potential expression

Equating the single-ion chemical potentials of the salt cation in both phases:

  • μ'_C(φ') = μ"_C(φ")

This yields:

  • φ' − φ" = (RT / z_C F) ln[(γ"_C m"_C) / (γ'_C m'_C)]

where:

  • φ is the electric potential
  • z_C is the charge of the cation
  • F is Faraday's constant
  • γ_C is the single-ion activity coefficient
  • m_C is the cation molality

📊 Example: negatively charged polyelectrolyte

Setup:

  • Phase ": aqueous solution of polyelectrolyte with net negative charge + counterion M⁺ + salt MX
  • Phase ': aqueous solution of salt MX only
  • Membrane: permeable to H₂O, M⁺, and X⁻; impermeable to polyelectrolyte

Initial state (nonequilibrium):

  • Phase " has more M⁺ ions than X⁻ ions (to balance the polyelectrolyte's negative charge).
  • Chemical potentials of both M⁺ and X⁻ are greater in phase " than in phase '.
  • Salt MX spontaneously passes from phase " to phase ' until equilibrium.

Equilibrium state (with approximations: pressure factors and activity coefficients ≈ 1):

  • For a 1:1 salt: (m'₊)(m'₋) ≈ (m"₊)(m"₋)
  • Electroneutrality conditions:
    • Phase ': m'₊ = m'₋
    • Phase ": m"₊ = m"₋ + |z_P|m_P (where z_P is the polyelectrolyte charge, m_P is its molality)
  • Substituting gives: (m'₋)² ≈ (m"₋ + |z_P|m_P)(m"₋)
  • This shows m'₋ > m"₋ and therefore m'₊ < m"₊ at equilibrium.

Electric potential:

  • For M⁺ to have the same chemical potential in both phases despite lower activity in phase ', the electric potential of phase ' must be greater than that of phase ".
  • φ' − φ" ≈ (RT/F) ln(m"₊ / m'₊) (positive in this example)
  • The Donnan potential results from a very small departure from exact electroneutrality: phase ' has a minute net positive charge, phase " has a net negative charge of equal magnitude.
  • The excess charge is distributed over the boundary surface, not in the bulk phase composition.

Pressure difference:

  • Can be estimated from: p" − p' ≈ (ρ_A RT) Σ(m"ᵢ − m'ᵢ) (summed over all solutes i ≠ A)
  • where ρ_A is the density of solvent A.
  • In this example, p" − p' is positive.
  • Without this pressure difference, solvent in phase ' would move spontaneously into phase " until phase ' disappears.

⚠️ Practical implications

  • The existence of a Donnan membrane equilibrium introduces complications.
  • It would make it difficult to:
    • Use a measured pressure difference to estimate the molar mass of the polyelectrolyte (by osmotic pressure methods).
    • Study the binding of a charged ligand by equilibrium dialysis.
  • Don't confuse: osmotic equilibrium (only pressure differs) vs. Donnan equilibrium (both pressure and electric potential differ, and ion distributions are unequal).

📐 Comparison of membrane equilibria

TypePermeable speciesImpermeable speciesPressure difference?Electric potential difference?Key application
OsmoticSolvent onlyAll solutesYesNoOsmotic pressure measurement
Equilibrium dialysisSolvent + small ligandMacromoleculeYesNo (uncharged ligand)Ligand-binding studies
DonnanSolvent + some ionsPolyelectrolyte + some ionsYesYesBiological membrane potentials
82

Liquid–Gas Equilibria

12.8 Liquid–Gas Equilibria

🧭 Overview

🧠 One-sentence thesis

When a liquid mixture is equilibrated with a gas phase, changes in liquid pressure or composition systematically affect gas-phase fugacities through thermodynamic relationships that predict solubility behavior, vapor pressure trends, and the formation of azeotropes.

📌 Key points (3–5)

  • Pressure effect on fugacity: Increasing the pressure of a liquid at constant composition slightly increases the fugacity of volatile components in the equilibrated gas phase (Poynting factor), though this effect is usually negligible unless the pressure change exceeds ~10 bar.
  • Composition coupling: In a binary liquid–gas system, a composition change that increases one component's gas fugacity must decrease the other's (Gibbs–Duhem constraint).
  • Raoult's and Henry's law connection: In the ideal-dilute region where the solute obeys Henry's law, the solvent must obey Raoult's law; if one component obeys Raoult's law at all compositions, so must the other (ideal mixture).
  • Common confusion—deviation patterns: If one component shows only positive deviations from Raoult's law (with one inflection point), the other component must also show positive deviations; the same applies to negative deviations.
  • Gas solubility principles: Gas solubility in liquids is proportional to partial pressure under ideal conditions, decreases with temperature (exothermic dissolution), and follows predictable trends based on solvent properties and Henry's law constants.

🔧 Effect of liquid pressure on gas fugacity

🔧 The Poynting factor

Poynting factor: The exponential term that relates the fugacity of a component in a gas phase equilibrated with a liquid at two different pressures.

  • At constant temperature and liquid composition, increasing the liquid pressure increases the fugacity of each volatile component in the gas phase.
  • The relationship is: f_i(p₂) = f_i(p₁) × exp[∫(V̄_i(l)/RT)dp from p₁ to p₂]
  • V̄_i(l) is the partial molar volume of component i in the liquid phase.

📏 When the effect matters

  • For typical partial molar volumes (~100 cm³/mol at 300 K):
    • Pressure change of 1 bar → fugacity ratio ~1.004 (0.4% change)
    • Pressure change of 10 bar → fugacity ratio ~1.04 (4% change)
    • Pressure change of 100 bar → fugacity ratio ~1.5 (50% change)
  • Key distinction: This applies only when the liquid phase is present and has constant composition; without the liquid, gas fugacity is approximately proportional to total pressure.

💧 Application to droplets

  • Small liquid droplets have slightly higher equilibrium vapor pressure than bulk liquid.
  • The smaller the droplet radius, the greater the fugacity and vapor pressure.
  • This is because the pressure inside a small droplet is elevated due to surface tension effects.

🔄 Effect of liquid composition on gas fugacities

🔄 The Gibbs–Duhem constraint

For a binary liquid mixture (A and B) equilibrated with a gas at constant T and p:

(x_A/f_A) df_A = −(x_B/f_B) df_B

  • This means: a composition change that increases component A's fugacity must decrease component B's fugacity.
  • The changes are coupled through the mole fractions and fugacities.

🧪 Deriving Raoult's law from Henry's law

Starting from Henry's law for the solute (f_B = k_H,B × x_B) in the ideal-dilute region:

  1. Henry's law gives: df_B/dx_B = k_H,B = f_B/x_B
  2. Combining with the Gibbs–Duhem constraint and integrating yields: f_A = x_A × f°_A
  3. This is Raoult's law for the solvent.

Two key conclusions:

  • In the ideal-dilute region, if the solute obeys Henry's law, the solvent must obey Raoult's law.
  • If one component of a binary mixture obeys Raoult's law at all compositions, the other must too (definition of an ideal mixture).

📊 Deviation patterns in nonideal mixtures

Deviation typeMathematical conditionConsequence for other component
Positive deviation (one inflection point)df_B/dx_B < f_B/x_BOther component must also show positive deviation
Negative deviation (one inflection point)df_B/dx_B > f_B/x_BOther component must also show negative deviation
Mixed deviationsMultiple inflection pointsUnusual behavior possible (e.g., chloroform–ethanol)

Example: Water–ethanol at 25°C shows positive deviations for both components, each with a single inflection point.

Don't confuse: The usual case (one inflection point) with unusual systems that have two inflection points, where one component can show positive deviations while the other shows both positive and negative deviations.

📐 The Duhem–Margules equation and Konowaloff's rule

📐 The Duhem–Margules equation

For a binary liquid–gas system at constant T and p:

(x_A/f_A)(df_A/dx_A) = −(x_B/f_B)(df_B/dx_A)

  • If the gas is ideal, fugacities equal partial pressures: (x_A/p_A)(dp_A/dx_A) = −(x_B/p_B)(dp_B/dx_A)
  • This relates the rate of change of each component's partial pressure to the liquid composition.

🌡️ Konowaloff's rule

Konowaloff's rule: Compared to the liquid phase, the gas phase is richer in the component whose addition to the liquid at constant temperature causes the total pressure to increase.

From the Duhem–Margules equation:

  • If y_A/y_B > x_A/x_B (gas richer in A than liquid), then dp/dx_A is positive (adding A increases total pressure).
  • If y_A/y_B < x_A/x_B (gas richer in B than liquid), then dp/dx_A is negative (adding A decreases total pressure).

🔁 Azeotropes

Azeotrope: A liquid composition at which the total pressure exhibits a maximum or minimum and the liquid and gas phases have identical mole fraction compositions.

  • At the azeotrope composition: x_A/x_B = y_A/y_B
  • At this point, dp/dx_A = 0 but dp_A/dx_A is still positive.
  • The liquid and gas cannot be separated by simple distillation at this composition.

🧊 Gas solubility

🧊 General solubility expression

For a nonelectrolyte gas B dissolving in a liquid:

x_B = (K × f_B/p°) / (γ_x,B × Π_x,B)

  • K is the thermodynamic equilibrium constant for B(g) → B(sln)
  • γ_x,B is the activity coefficient (mole fraction basis)
  • Π_x,B is the pressure factor
  • At fixed T and p, increasing the activity coefficient decreases solubility for the same gas fugacity.

🧂 The salting-out effect

  • Dissolving a salt in an aqueous solution often increases the activity coefficient of a dissolved gas.
  • This decreases the gas solubility—the "salting-out effect."
  • Example: Adding NaCl to water decreases the solubility of dissolved O₂.

📉 Simplified solubility law

Under ideal conditions (low pressure, dilute solution, ideal gas):

x_B = K × (p_B/p°)

  • Solubility is proportional to partial pressure (Henry's law in pressure form).
  • This is valid when Π_x,B ≈ 1 and γ_x,B ≈ 1.

🌡️ Temperature dependence

The temperature dependence of gas solubility at fixed partial pressure:

∂(ln x_B)/∂T = ΔH°_sol,B / (RT²)

  • Since gas dissolution is exothermic (ΔH°_sol,B < 0), solubility decreases with increasing temperature.
  • A plot of ln x_B versus 1/T gives a straight line with slope −ΔH°_sol,B/R.

🎯 Ideal solubility predictions

Ideal solubility of a gas: x_B = p_B/p_B, where p_B is the vapor pressure of pure liquid B.

Four predictions for ideal solubility at fixed p_B:

  1. Solvent independence: Ideal solubility (as mole fraction) is independent of solvent type.
  2. Concentration effect: Solubility as concentration (c_B) is lower for solvents with larger molar volumes.
  3. Volatility effect: More volatile pure liquid solutes (higher p*_B) have lower solubility.
  4. Temperature effect: Solubility decreases with increasing temperature (since p*_B increases).

Example: Cl₂ solubility at 0°C and 1.01 bar is x_B = 0.270 in heptane, 0.288 in SiCl₄, and 0.298 in CCl₄—all close to the ideal value of 0.273.

🔬 Henry's law constants: temperature and pressure effects

🔬 Relationship to equilibrium constant

For a nonelectrolyte solute with mole-fraction standard state:

k_H,B = (Π_x,B × p°) / K

  • k_H,B depends on both temperature and pressure.
  • At standard pressure (p° = 1 bar), Π_x,B = 1.

🌡️ Temperature dependence at standard pressure

d ln(k_H,B/p°)/dT = −ΔH°_sol,B / (RT²)

or equivalently:

d ln(k_H,B/p°)/d(1/T) = ΔH°_sol,B / R

  • A plot of ln(k_H,B/p°) versus 1/T yields ΔH°_sol,B from the slope.
  • These expressions are valid up to ~2 bar with negligible error.

📊 Pressure dependence

At constant temperature, comparing two pressures p₁ and p₂:

k_H,B(p₂) = k_H,B(p₁) × exp[∫(V̄°_B/RT)dp from p₁ to p₂]

  • V̄°_B is the standard partial molar volume of the solute.
  • If V̄°_B is approximately constant: k_H,B(p₂) ≈ k_H,B(p₁) × exp[V̄°_B(p₂ − p₁)/(RT)]
  • Unless |p₂ − p₁| is much greater than 1 bar, the pressure effect on k_H,B is small.

Don't confuse: The pressure effect on k_H,B (usually small) with the direct proportionality of solubility to partial pressure (Henry's law itself).

83

Reaction Equilibria

12.9 Reaction Equilibria

🧭 Overview

🧠 One-sentence thesis

The thermodynamic equilibrium constant for a reaction can be expressed using activities based on appropriate standard states for each species, and changes in conditions like ionic strength alter the equilibrium composition to maintain the constant K value.

📌 Key points (3–5)

  • Definition of K: the equilibrium constant is the product of activities of all species, each raised to its stoichiometric coefficient.
  • Activities depend on standard states: each species (gas, pure solid/liquid, solute, ion) uses a different activity expression from Table 12.2.
  • Mixed equilibrium constants: reactions involving multiple phases use more than one kind of standard state in the same K expression.
  • Common confusion: K depends only on temperature, but activity coefficients (like γ±) depend on ionic strength—so changing ionic strength forces molalities to adjust to keep K constant.
  • Practical implication: adding an inert salt changes ionic strength, which changes activity coefficients and therefore shifts the equilibrium composition (e.g., degree of dissociation).

🧮 The equilibrium constant expression

🧮 Definition and structure

The thermodynamic equilibrium constant: K = product of (activity of species i)^(stoichiometric coefficient νᵢ) for all reactants and products at equilibrium.

  • Written as K = ∏ (aᵢ)^νᵢ (Eq. 12.9.1).
  • Each activity aᵢ is based on an appropriate standard state for that species.
  • The value of K depends only on temperature T, not on pressure or composition.

🔄 Replacing activities with measurable quantities

  • Table 12.2 provides expressions for activities of different types of species:
    • Pure gas: activity = fugacity / standard pressure
    • Pure solid or liquid: activity = pressure factor γ
    • Gas in a mixture: activity = fugacity of component / standard pressure
    • Solute (nonelectrolyte, molality basis): activity = (activity coefficient) × (molality / standard molality)
    • Ion in solution: activity = (activity coefficient) × (molality / standard molality)
    • Electrolyte solute: activity involves mean ionic activity coefficient γ± and molalities of ions
  • By substituting these expressions into the equilibrium constant, we obtain a working equation with measurable quantities.

🪨 Heterogeneous equilibrium example

🪨 Limestone cavern reaction

The excerpt gives the reaction: CaCO₃(cr, calcite) + CO₂(g) + H₂O(sln) ⇌ Ca²⁺(aq) + 2HCO₃⁻(aq)

  • Treating H₂O as solvent and Ca²⁺, HCO₃⁻ as solutes.
  • The equilibrium constant is written as: K = (a₊ · a₋²) / (a_CaCO₃ · a_CO₂ · a_H₂O)
  • Substituting activity expressions (Eq. 12.9.2): K = [γ₊ · γ₋² · (m₊ · m₋² / (m°)³)] / [(f_CO₂ / p°) · (γ_H₂O · x_H₂O)]
  • The term Γᵣ is a quotient of pressure factors for condensed phases; at low pressure, Γᵣ ≈ 1.

🧂 Mean ionic activity coefficient

  • For the electrolyte Ca(HCO₃)₂, the product γ₊ · γ₋² can be replaced by γ±³, where γ± is the mean ionic activity coefficient.
  • Equation 12.9.4: K = [Γᵣ · γ±³ · (m₊ · m₋² / (m°)³)] / [(f_CO₂ / p°) · (γ_H₂O · x_H₂O)]
  • Alternatively, treat dissolved Ca(HCO₃)₂ as a single solute species B (Eq. 12.9.5), then substitute the electrolyte activity expression from Table 12.2 to recover the same form.

🧪 Effect of adding inert salt

  • K depends only on T; Γᵣ depends only on T and p.
  • Example: dissolving NaCl at constant T and p increases ionic strength.
  • This changes γ±, so the solute molalities must adjust to keep K constant—the system shifts to a new equilibrium composition.
  • Don't confuse: the equilibrium constant itself does not change; only the distribution of species changes to maintain K.

🍋 Weak acid dissociation example

🍋 Acetic acid ionization

Reaction: HA(aq) ⇌ H⁺(aq) + A⁻(aq)

  • The acid dissociation constant Kₐ is defined (Eq. 12.9.6): Kₐ = [Γᵣ · γ₊ · γ₋ · (m₊ · m₋ / m_HA)] / (γ_m,HA · m°)
  • Using the mean ionic activity coefficient: Kₐ = [Γᵣ · γ±² · (m₊ · m₋ / m_HA)] / (γ_m,HA · m°)

🔢 Degree of dissociation

  • Assume the solution is prepared from water and acid, and H⁺ from water dissociation is negligible.
  • Let α = degree of dissociation, m_B = overall molality of acid.
  • Then m₊ = m₋ = α · m_B and m_HA = (1 − α) · m_B.
  • Equation 12.9.7: Kₐ = [Γᵣ · γ±² · α² · m_B / m°] / [γ_m,HA · (1 − α)]

📈 Ionic strength effect on dissociation

  • At constant T, p, and m_B, if ionic strength increases (e.g., by adding an inert salt), γ± decreases.
  • To keep Kₐ constant, the degree of dissociation α must increase.
  • Example: adding NaCl to an acetic acid solution lowers γ± and shifts the equilibrium toward more dissociation.
  • Don't confuse: the dissociation constant Kₐ itself is fixed by temperature; only the equilibrium composition (α) changes.

📐 Standard molar quantities

📐 Relationship among standard quantities

Standard molar reaction Gibbs energy: ΔᵣG° = ΔᵣH° − T·ΔᵣS° (Eq. 12.10.1)

  • ΔᵣG°, ΔᵣH°, and ΔᵣS° are standard molar reaction Gibbs energy, enthalpy, and entropy.
  • These are the most useful experimentally-derived data for thermodynamic calculations.

📐 Standard molar reaction entropy

  • Calculated from standard molar entropies of reactants and products (Eq. 12.10.2): ΔᵣS° = Σ νᵢ · S°ᵢ
  • The sum runs over all species i, with stoichiometric coefficients νᵢ.

🔬 Experimental evaluation methods

The excerpt lists techniques for obtaining these quantities:

QuantityMethodNotes
ΔᵣH°Reaction calorimetryMeasures heat of reaction (Sec. 11.5)
S°ᵢ (solid or liquid)CalorimetryHeat capacity and phase-transition enthalpy measurements (Sec. 6.2.1)
S°ᵢ (gas)SpectroscopySpectroscopic measurements (Sec. 6.2.2)
  • These experimental values allow calculation of ΔᵣG° via Eq. 12.10.1, which is related to the equilibrium constant K.
84

Evaluation of Standard Molar Quantities

12.10 Evaluation of Standard Molar Quantities

🧭 Overview

🧠 One-sentence thesis

Standard molar reaction quantities (enthalpy, Gibbs energy, and entropy) can be evaluated through multiple experimental techniques and efficiently tabulated as formation values to calculate reaction properties for reactions not directly investigated.

📌 Key points (3–5)

  • What is evaluated: standard molar reaction enthalpies (ΔrH°), Gibbs energies (ΔrG°), and entropies (ΔrS°) for reactions.
  • How they relate: these three quantities are connected by ΔrG° = ΔrH° − TΔrS°, allowing calculation of one from the others.
  • Experimental methods: calorimetry, spectroscopy, equilibrium constant measurements, and cell potential measurements all provide pathways to these values.
  • Efficient tabulation: formation values (ΔfH° and ΔfG°) allow calculation of reaction quantities for any reaction from a single database.
  • Common confusion: standard molar reaction quantities vs. standard molar formation quantities—formation values are for making a compound from elements, while reaction values are for any specified reaction.

🔗 Relationships among standard quantities

🔗 The fundamental equation

ΔrG° = ΔrH° − TΔrS°

  • This equation (12.10.1) connects the three main standard molar reaction quantities.
  • If you know any two of these values at a given temperature, you can calculate the third.
  • The entropy term includes temperature explicitly, showing how Gibbs energy changes with temperature even when enthalpy and entropy are constant.

🔗 Calculating reaction entropy

ΔrS° = Σᵢ νᵢSᵢ°

  • Equation 12.10.2 shows that the standard molar reaction entropy is the sum of standard molar entropies of products minus reactants, weighted by stoichiometric coefficients νᵢ.
  • This requires knowing the standard molar entropy Sᵢ° for each substance involved.
  • Example: for a reaction with 2 moles of product A and 1 mole of reactant B, ΔrS° = 2S°(A) − S°(B).

🧪 Experimental techniques for evaluation

🧪 Calorimetry methods

  • Reaction calorimetry directly measures ΔrH° for a reaction (referenced to Section 11.5).
  • Heat capacity and phase-transition measurements provide S°ᵢ for solids and liquids (Section 6.2.1).
  • These methods involve direct heat measurements under controlled conditions.

🔬 Spectroscopy

  • Spectroscopic measurements can evaluate S°ᵢ for gases (Section 6.2.2).
  • This provides an alternative to calorimetric methods for gaseous substances.

⚖️ Equilibrium constant measurements

  • Measuring the thermodynamic equilibrium constant K and its temperature dependence provides both ΔrG° and ΔrH°.
  • The relations used are:
    • ΔrG° = −RT ln K
    • ΔrH° = R d(ln K)/d(1/T)
  • This works for any equilibrium: vapor pressure, solubility, chemical reaction, etc.
  • Don't confuse: the temperature derivative is with respect to 1/T, not T directly.

🔋 Electrochemical methods

  • For reactions that can run reversibly in a galvanic cell, cell potential measurements are useful.
  • The standard cell potential and its temperature derivative allow evaluation of ΔrH°, ΔrG°, and ΔrS° (details in Section 14.3.3).

📊 Formation quantities as a tabulation system

📊 Why formation values are efficient

Standard molar enthalpies and Gibbs energies of formation: values for making a compound from its elements in their standard states.

  • Instead of tabulating reaction quantities for every possible reaction, we tabulate formation values for each substance.
  • Any reaction's standard quantities can then be calculated from formation values using equations 12.10.3:
    • ΔrH° = Σᵢ νᵢΔfH°(i)
    • ΔrG° = Σᵢ νᵢΔfG°(i)
  • Example: to find ΔrH° for A + B → C, calculate ΔfH°(C) − ΔfH°(A) − ΔfH°(B).

📊 Conventions for ions

The excerpt establishes reference conventions for aqueous hydrogen ions:

  • ΔfH°(H⁺, aq) = 0
  • ΔfG°(H⁺, aq) = 0
  • S°ₘ(H⁺, aq) = 0

These conventions allow consistent tabulation of properties for all other ions relative to H⁺.

📊 Available data

  • Appendix H provides an abbreviated set of ΔfH°, S°ₘ, and ΔfG° values at 298.15 K.
  • These values come from the various experimental techniques described above.
  • Problems 12.18–12.20, 14.3, and 14.4 provide examples of evaluating these quantities from experimental measurements.

🔄 Practical workflow

🔄 From measurements to reaction quantities

Starting pointWhat you measureWhat you calculateMethod
Direct reactionHeat released/absorbedΔrH°Reaction calorimetry
Equilibrium systemK at multiple temperaturesΔrG° and ΔrH°ln K vs. 1/T plot
Pure substanceHeat capacity, phase transitionsS°ᵢCalorimetry (solid/liquid) or spectroscopy (gas)
Galvanic cellCell potential and temperature dependenceΔrH°, ΔrG°, ΔrS°Electrochemistry

🔄 From formation values to any reaction

  • Look up ΔfH° and ΔfG° for all reactants and products in tables.
  • Apply the summation formulas (equations 12.10.3) with stoichiometric coefficients.
  • This allows calculation of standard quantities for reactions not directly investigated.
  • Don't confuse: formation quantities are always for making the substance from elements; reaction quantities are for the specific reaction of interest.
85

The Gibbs Phase Rule for Multicomponent Systems

13.1 The Gibbs Phase Rule for Multicomponent Systems

🧭 Overview

🧠 One-sentence thesis

The Gibbs phase rule determines how many intensive variables (degrees of freedom) can be independently varied in a multicomponent equilibrium system without changing the number or kinds of phases present.

📌 Key points

  • What the phase rule calculates: the number of degrees of freedom (F), which is the maximum number of intensive variables that can be varied independently while maintaining the same phases and species in equilibrium.
  • Two equivalent formulations: F = 2 + C − P (components approach) or F = 2 + s − r − P (species approach), where C = s − r.
  • Key distinction—species vs. components: species are distinct chemical entities (e.g., CO₂, CO₃²⁻), while components are the minimum number of substances needed to prepare each phase individually.
  • Common confusion—what counts as r: r includes independent reaction equilibria, electroneutrality requirements for phases with ions, and initial conditions that create relations among intensive variables (not extensive ones).
  • Why it matters: the phase rule predicts how many intensive properties (like T, p, mole fractions) can be freely chosen before the system's equilibrium state is fully determined.

🔢 Understanding degrees of freedom

🔢 What F represents

The number of degrees of freedom (or variance), F: the maximum number of intensive variables that can be varied independently while the system remains in an equilibrium state.

  • F is not about how much of each phase is present (extensive properties); it is about which intensive properties (T, p, composition) can be independently chosen.
  • "Independently varied" means you can set arbitrary values within a certain range without causing a phase to appear or disappear.
  • Example: If F = 2, you might independently choose T and p; then all other intensive variables (like mole fractions) are determined by the system's nature.

🔄 What changes are allowed

  • The phase rule assumes the system remains in thermal, mechanical, and transfer equilibrium as you vary intensive properties.
  • Phases are not separated by adiabatic/rigid partitions or semipermeable membranes.
  • Every conceivable reaction is either at equilibrium or frozen at a fixed advancement during observation.
  • Changing the amount of a phase (an extensive variable) does not count as a degree of freedom.

🚫 Don't confuse: intensive vs. extensive constraints

  • If an initial condition creates a relation among amounts of phases (extensive variables), it does not reduce F.
  • Only relations among intensive variables (like chemical potentials, mole fractions) reduce the number of degrees of freedom.
  • Example: In the carbon-oxygen system, the relation n_C = 2n_O₂ + n_CO₂ is extensive and does not change r or F.

🧬 Species approach to the phase rule

🧬 Counting variables and constraints

A species is an entity distinguished by its chemical formula (e.g., CO₂(aq) and CO₂(g) are the same species in different phases).

  • Start with 2 + P·s independent variables: T, p, and the amount of each of s species in each of P phases.
  • Transfer equilibrium imposes s(P − 1) constraints: each species must have the same chemical potential in all phases (μ_i^α = μ_i^β = ...).
  • After accounting for transfer equilibrium: 2 + P·s − s(P − 1) = 2 + s independent variables.

🔗 Additional constraints (r)

The term r counts independent relations among intensive variables beyond thermal, mechanical, and transfer equilibrium:

Source of constraintWhat it contributesExample
Reaction equilibriaEach independent equilibrium adds 1 to rΔ_r G = Σ ν_i μ_i = 0
ElectroneutralityEach phase with ions adds 1 to r2m_Pb²⁺ = m_Cl⁻ in aqueous phase
Initial conditionsOnly if they create intensive-variable relationsFixed mole ratio in gas phase
  • The final formula: F = 2 + s − r − P
  • If a species is absent from a phase, there is one fewer amount variable and one fewer transfer equilibrium relation; these cancel, so F is unchanged.

📐 Example: liquid water with ion equilibrium

  • Species: H₂O, H₃O⁺, OH⁻ → s = 3
  • Relations: (1) reaction equilibrium 2H₂O ⇌ H₃O⁺ + OH⁻, (2) electroneutrality m_H₃O⁺ = m_OH⁻ → r = 2
  • Phases: 1 liquid phase → P = 1
  • Result: F = 2 + 3 − 2 − 1 = 2
  • Don't confuse: even if you consider more ion species (e.g., H₅O₂⁺), each adds 1 to s and 1 to r, so F remains 2.

🧩 Components approach to the phase rule

🧩 What components are

The number of components, C: the minimum number of substances (or fixed-composition mixtures) from which each individual phase could be prepared, using methods that may be hypothetical.

  • Preparation methods may include adding/removing substances or using reactions that are at equilibrium in the actual system.
  • Nothing must remain unused when preparing a phase.
  • C may be less than the number of substances present if reactions produce some substances from others.

🔍 How to count components

  • Each substance present is either a component or can be formed from components by a reaction at equilibrium.
  • Example: CaCO₃(s), CaO(s), CO₂(g) system with equilibrium CaCO₃ ⇌ CaO + CO₂:
    • Cannot prepare CaO phase from CaCO₃ alone (CO₂ would be left over).
    • Can prepare CaCO₃ from CaO + CO₂ by reverse reaction.
    • Components: CaO and CO₂ → C = 2.

🧮 Deriving F from components

  • Start with 2 + P(C − 1) intensive variables: T, p, and (C − 1) mole fractions per phase.
  • Transfer equilibrium for each component: C(P − 1) constraints.
  • Result: F = [2 + P(C − 1)] − C(P − 1) = 2 + C − P
  • If a component is absent from a phase or a substance is formed by reaction, the changes in variables and constraints cancel.

🔗 Relationship between approaches

The two formulations are equivalent:

  • C = s − r
  • Components approach: F = 2 + C − P
  • Species approach: F = 2 + s − r − P
  • Both give the same F because electroneutrality and initial conditions are implicit in component definitions.

🔬 Worked examples

🔬 Pure liquid water

  • Components approach: C = 1 (H₂O), P = 1 → F = 2 + 1 − 1 = 2
  • Species approach: s = 1 (H₂O), r = 0, P = 1 → F = 2 + 1 − 0 − 1 = 2
  • Interpretation: can independently vary two intensive properties, e.g., T and p, or T and ρ.

🔬 Carbon, oxygen, and carbon oxides with reactions

  • System: solid C(s) and gas phase with O₂, CO, CO₂
  • Reactions at equilibrium: (1) 2C + O₂ ⇌ 2CO, (2) C + O₂ ⇌ CO₂
  • Species approach: s = 4, r = 2 (two independent equilibria), P = 2 → F = 2 + 4 − 2 − 2 = 2
  • Components approach: can prepare solid from C and gas from C + O₂ using both reactions → C = 2, P = 2 → F = 2 + 2 − 2 = 2
  • Interpretation: can arbitrarily vary T and p; then gas-phase mole fractions are determined by the equilibria.
  • Initial condition check: if prepared with n_C = 2n_O₂ initially, the relation n_C = 2n_O₂ + n_CO₂ is extensive only, so r and F are unchanged.

🔬 Solid salt and saturated solution

  • System: PbCl₂(s) and aqueous phase with H₂O, Pb²⁺(aq), Cl⁻(aq)
  • Components approach: prepare solid from PbCl₂, aqueous from PbCl₂ + H₂O → C = 2, P = 2 → F = 2
  • Species approach: s = 4 (PbCl₂, Pb²⁺, Cl⁻, H₂O), r = 2 (dissolution equilibrium + electroneutrality 2m_Pb²⁺ = m_Cl⁻), P = 2 → F = 2 + 4 − 2 − 2 = 2

🔬 Water and water-saturated air

  • System: liquid water with dissolved N₂, O₂ and gas phase with H₂O, N₂, O₂
  • No special initial condition: C = 3 (H₂O, N₂, O₂), P = 2 → F = 2 + 3 − 2 = 3
  • Can independently specify T, p, and y_N₂; then y_H₂O and x_N₂ are determined.
  • With fixed N₂/O₂ ratio in initial dry air: the relation (n_N₂^l + n_N₂^g)/(n_O₂^l + n_O₂^g) = a is extensive only, so r = 0 and F = 3 still.
  • Don't confuse: the mole ratio in solution may differ from that in gas, so air does not behave as a single substance; all three components are still needed.

📋 Summary table

ApproachFormulaKey terms
ComponentsF = 2 + C − PC = minimum substances to prepare each phase
SpeciesF = 2 + s − r − Ps = number of distinct chemical entities; r = independent intensive constraints
RelationC = s − rBoth formulas are equivalent

What counts in r:

  • Independent reaction equilibria (Δ_r G = 0)
  • Electroneutrality of phases with ions
  • Initial conditions that create intensive-variable relations (not extensive-only relations)
86

Phase Diagrams: Binary Systems

13.2 Phase Diagrams: Binary Systems

🧭 Overview

🧠 One-sentence thesis

Binary phase diagrams map the stable phases of two-component systems as functions of temperature and composition (or pressure and composition), revealing how many phases coexist, their compositions, and phenomena like eutectic points, azeotropes, and retrograde condensation.

📌 Key points (3–5)

  • Degrees of freedom: A binary system has C = 2, so F = 4 − P; maximum 3 degrees of freedom (one phase) and maximum 4 phases.
  • Two-dimensional diagrams: Usually plot temperature–composition at fixed pressure or pressure–composition at fixed temperature; the system point's location determines phase behavior.
  • Reading tie lines: In two-phase areas, horizontal tie lines connect the compositions of coexisting phases; the lever rule gives relative amounts.
  • Common confusion: Liquidus vs vaporus curves—liquidus shows liquid composition in equilibrium with another phase; vaporus shows gas composition; they coincide at azeotropic points but not elsewhere.
  • Special points and phenomena: Eutectic points (three phases, lowest melting temperature), azeotropes (liquid and gas have identical composition), and retrograde condensation (compression causes condensation then re-vaporization).

📐 Reading binary phase diagrams

📐 One-phase regions

One-phase area: The composition variable directly gives the composition of that single phase.

  • Three degrees of freedom in a binary system with one phase.
  • On a T–composition diagram at fixed pressure, you can independently vary T and composition within the one-phase boundary.
  • Example: A point in the "liquid" area means the entire system is a single liquid mixture of that composition at that temperature.

🔗 Two-phase regions and tie lines

Tie line: A horizontal line of constant T (on T–composition diagrams) or constant p (on p–composition diagrams) through a two-phase area.

  • The ends of the tie line give the compositions of the two coexisting phases.
  • The lever rule applies: the ratio of amounts in the two phases equals the ratio of distances from the system point to the tie-line ends.
  • Example: If the system point is at overall composition z_B = 0.40 in a two-phase area, draw a horizontal tie line; the left end might show x'_B = 0.20 (phase α composition) and the right end x''_B = 0.92 (phase β composition). The lever rule gives n''/n' = (0.40 − 0.20)/(0.92 − 0.40) = 0.38.

🔺 Three-phase equilibria

  • A binary system with three phases has F = 1 (univariant).
  • Cannot be represented by an area; appears as a horizontal line on 2D diagrams.
  • The three phase compositions are given by special points: the two ends of the line plus a junction point.
  • The system point's position along this line does not uniquely determine relative amounts of the three phases.

🗺️ Boundary curves: liquidus, solidus, vaporus

Curve nameWhat it shows
LiquidusComposition of liquid phase in equilibrium with another phase (also called bubble-point or boiling-point curve)
SolidusComposition of solid phase in equilibrium with another phase
VaporusComposition of gas phase in equilibrium with another phase (also called dew-point or condensation curve)
  • These curves separate one-phase areas from two-phase areas.
  • Don't confuse: The liquidus curve is not "where liquid exists"—it's the boundary showing liquid composition when two phases coexist.

🧊 Solid–liquid systems

❄️ Eutectic behavior

Eutectic point: The point where two solid phases and one liquid phase coexist; the eutectic temperature T_e is the lowest temperature at which liquid is stable at that pressure.

  • The liquidus curves from two pure components meet at the eutectic point.
  • Eutectic composition: the liquid composition at this point.
  • Eutectic halt: During cooling, temperature stops decreasing while the eutectic liquid freezes into two solid phases.
  • Example: Cooling a liquid from above the eutectic temperature—solid A begins to freeze out first, then at T_e both solid A and solid B freeze simultaneously until all liquid is gone.

🧪 Thermal analysis

  • Experimental method: observe cooling curves (temperature vs time) at various compositions.
  • A break (change in slope) indicates entering a two-phase region.
  • A halt (constant temperature) indicates either a pure-component freezing point or a eutectic temperature.

🏗️ Solid solutions vs pure solids

  • Some systems form pure solid crystals (e.g., chloroform–carbon tetrachloride).
  • Others form solid solutions (substitutional alloys) with variable composition (e.g., silver–copper system with phases s' and s'').
  • In solid-solution systems, tie lines in two-phase areas do not end at vertical lines (pure components) but at curved boundaries.

💎 Solid compounds

Congruent melting (dystectic reaction): A solid compound melts to give a liquid of the same composition.

Incongruent melting (peritectic reaction): A solid compound decomposes into a liquid and a different solid.

  • Example of congruent: α-naphthylamine–phenol system with compound AB melting at its own composition.
  • The liquidus curve is rounded (not a cusp) at the compound composition because the compound dissociates in the liquid.
  • Peritectic point: Where a solid compound decomposes; three phases coexist (e.g., NaCl·2H₂O → solution + NaCl(s) at 0 °C).

💧 Liquid–liquid systems

🌊 Partially miscible liquids

Conjugate phases: Two liquid phases in equilibrium with one another.

Miscibility gap: The difference in compositions between two conjugate liquid phases.

  • At a given T and p, each conjugate phase has a fixed composition (F = 2).
  • The miscibility gap typically decreases as temperature increases.

🔥 Upper consolute temperature

Upper consolute temperature (upper critical solution temperature): The temperature above which the miscibility gap vanishes and the system forms a single liquid phase.

Consolute point (critical point): The point at the maximum of the two-phase boundary where the two liquid phases become identical.

  • Critical opalescence occurs near this point due to large composition fluctuations.
  • Example: Methyl acetate–carbon disulfide system—at low T two liquid layers exist; heating causes the layers to merge into one phase.

🔄 Phase changes with composition or temperature

  • Adding more of one component at constant T: system point moves along tie line; phase compositions stay fixed but relative amounts change.
  • Heating at constant composition: system point moves up an isopleth (vertical line); eventually enters one-phase region when miscibility gap is crossed.

🌡️ Liquid–gas systems: ideal mixtures

📏 Raoult's law systems

For ideal liquid mixtures obeying Raoult's law (p_A = x_A p*_A):

  • Total pressure: p = p_B + (p_A − p*_B)x_A (linear in liquid composition x_A)
  • Gas composition: y_A = x_A p*_A / p
  • At constant T, pressure is linear with liquid composition but not with gas composition.

📊 Generating phase diagrams

Given vapor pressures p_A(T) and p_B(T):

  • Liquid composition at given T and p: x_A = (p − p_B)/(p_A − p*_B)
  • Gas composition: y_A = x_A p*_A / p
  • These equations generate liquidus and vaporus curves.
  • Example: Toluene–benzene system shows linear liquidus on p–composition diagram but curved liquidus on T–composition diagram.

🗻 Three-dimensional view

  • Liquidus and vaporus are actually surfaces in T–p–z_A space.
  • Two-dimensional diagrams are cross-sections at constant T or constant p.
  • Tie lines connect points on the two surfaces at the same T and p.

🌀 Liquid–gas systems: nonideal mixtures

⚡ Deviations from Raoult's law

  • Positive deviations: partial pressures higher than Raoult's law predicts; molecules prefer to escape the liquid.
  • Negative deviations: partial pressures lower than predicted; specific A–B interactions (solvation, association) favor mixing.
  • Most mixtures show positive deviations.

🎯 Azeotropic behavior

Azeotrope (azeotropic mixture): A liquid mixture that vaporizes to give a gas of identical composition; boils unchanged.

Azeotropic point: Where liquidus and vaporus curves coincide on a phase diagram.

  • Occurs when deviations from Raoult's law are large enough.
  • Positive deviations → pressure maximum on p–composition diagram → minimum-boiling azeotrope on T–composition diagram.
  • Negative deviations → pressure minimum → maximum-boiling azeotrope.
  • Example: Methanol–benzene at 45 °C shows positive deviations; azeotrope at z_A = 0.59 and p = 60.5 kPa.

🔬 Experimental determination

  • Use equilibrium still: boil liquid mixtures at fixed T or p.
  • Sample both phases after equilibration and analyze compositions.
  • Calculate partial pressures from gas composition: p_A = y_A p.

🧮 Degrees of freedom at azeotrope

  • Two species, two phases, one relation (x_A = y_A).
  • F = 2 + s − r − P = 2 + 2 − 1 − 2 = 1 (univariant).
  • At given T, azeotrope exists at only one p and one composition.
  • Azeotrope vapor-pressure curve: shows how azeotrope composition and pressure vary with T.

🚫 Fractional distillation limitation

  • Zeotropic systems (no azeotrope): can separate into pure components by fractional distillation.
  • Azeotropic systems: distillation separates into one pure component and the azeotropic mixture, not two pure components.

🌈 Complex liquid–gas diagrams

  • Partially miscible liquids that boil before reaching upper consolute temperature show combined liquid–liquid and liquid–gas equilibria.
  • Can exhibit minimum-boiling azeotropes or boiling below either pure component's boiling point.

⚗️ Solid–gas systems

💨 Hydrate dissociation

Example: CuSO₄–H₂O system with hydrates CuSO₄·H₂O, CuSO₄·3H₂O, CuSO₄·5H₂O.

Dissociation pressure: The unique pressure at which a hydrate, its dehydration product, and water vapor coexist at a given temperature.

  • Dissociation equilibrium example: CuSO₄·5H₂O(s) ⇌ ½CuSO₄·3H₂O(s) + H₂O(g)
  • Three phases, two components → F = 1 (univariant).
  • Appears as horizontal line on p–composition diagram at fixed T.
  • The equilibrium constant approximately equals p_d/p° (dissociation pressure over standard pressure).

🌬️ Efflorescence and deliquescence

When dry air is present:

Efflorescence: Spontaneous dehydration when partial pressure of H₂O in air is less than the dissociation pressure of the hydrate.

Deliquescence: Spontaneous absorption of water to form saturated solution when partial pressure of H₂O exceeds the vapor pressure of the saturated solution.

  • Below dissociation pressure: hydrate cannot exist in equilibrium with water vapor.
  • Above dissociation pressure: dehydrated salt cannot exist in equilibrium with water vapor.

🚀 High-pressure systems

🌊 Critical curves

Critical curve: The locus of critical points where gas and liquid mixtures become identical in composition and density.

  • Extends from the critical point of one pure component toward (or beyond) the critical point of the other.
  • Two types observed:
    • Type 1 (heptane–ethane): critical curve runs between the two pure-component critical points.
    • Type 2 (xenon–helium): critical curve extends to higher T and p than either pure-component critical point.

🔄 Retrograde phenomena

Retrograde condensation: Isothermal compression causes condensation followed by re-vaporization.

Retrograde vaporization: Isobaric heating causes vaporization followed by re-condensation.

  • Example: In heptane–ethane system at 450 K, increasing pressure along a certain path causes the system to enter then exit the two-phase region.
  • Occurs because the critical curve creates unusual phase-boundary shapes.

🌫️ Gas–gas immiscibility

  • At pressures above the critical curve in Type 2 systems, two high-density fluid phases coexist.
  • Sometimes called "gas–gas equilibrium" even though both phases have liquid-like densities.
  • Demonstrates that high-pressure gases do not necessarily mix spontaneously.
87

Phase Diagrams: Ternary Systems

13.3 Phase Diagrams: Ternary Systems

🧭 Overview

🧠 One-sentence thesis

Ternary phase diagrams use triangular coordinates to represent three-component systems at constant temperature and pressure, showing how composition determines the number and identity of phases present.

📌 Key points (3–5)

  • What ternary systems are: systems with three components where temperature, pressure, and two independent composition variables can be varied.
  • How triangular coordinates work: each vertex represents a pure component; distance from a side to a point (as a fraction of triangle height) gives the mole fraction of the component at the opposite vertex.
  • Two-phase regions: tie lines connect the compositions of coexisting phases; the lever rule determines relative amounts.
  • Common confusion: the sum of perpendicular distances from any point to the three sides always equals the triangle height, which is why mole fractions always sum to 1.
  • Why it matters: ternary diagrams reveal miscibility gaps, plait points, eutonic points, and phenomena like the common ion effect.

📐 Triangular coordinate system

📐 How to read the triangle

A point within an equilateral triangle represents a ternary system with all three components; each vertex represents one pure component, and each side represents a binary system of the two components at the ends of that side.

  • To find the mole fraction z_A of component A: measure the perpendicular distance from the point to the side opposite vertex A, then divide by the triangle height.
  • The same procedure applies to z_B and z_C.
  • The sum z_A + z_B + z_C must equal 1.

Why this works: The sum of the perpendicular lines drawn from any point to the three sides equals the height of the triangle (proof given in the excerpt using nested equilateral triangles).

📏 Grid lines and reading compositions

  • Equally-spaced lines parallel to each side help convert position to composition.
  • Each line parallel to a side represents a constant mole fraction of one component.
  • Example from the excerpt: lines dividing the distance into ten equal parts represent mole fraction differences of 0.1, starting at 0 at the side and ending at 1 at the opposite vertex.
  • A filled circle at z_A = 0.20, z_B = 0.30, z_C = 0.50 can be read directly using these grid lines.

🔍 Two useful properties

PropertyWhat it meansExample
Lines parallel to a sideOne mole fraction remains constantMoving along such a line changes only the ratio of the other two components
Lines through a vertexThe ratio of two mole fractions remains constantMoving toward a vertex increases that component while keeping the ratio of the other two fixed

🧪 Three-liquid system: ethanol–benzene–water

🧪 Single-phase and two-phase regions

  • P = 1 region: a single liquid phase whose composition is given by the point position.
  • The one-phase area extends to the ethanol–benzene side and the ethanol–water side, meaning these pairs mix in all proportions.
  • P = 2 region: two liquid phases coexist.
  • The compositions of the two phases are given by the ends of a tie line through the system point.
  • Tie lines must be determined experimentally; four representative tie lines are shown in the diagram.

⚖️ Lever rule and relative amounts

  • The relative amounts of the two phases can be determined from the lever rule.
  • The lever rule works because the ratio n_A / n (equal to z_A) varies linearly with the position of the system point along a tie line.
  • Don't confuse: the system point is not the composition of either phase; it is the overall composition, and the tie line endpoints give the phase compositions.

🎯 Plait point (critical solution point)

The plait point is the composition at which the miscibility gap disappears and the compositions of two conjugate liquid phases become identical.

  • As the system point approaches the plait point from within the two-phase area, the tie line length approaches zero.
  • At the plait point, the two-phase region ends.

💧 Adding ethanol to benzene–water

Example scenario from the excerpt:

  • Point a: binary benzene–water system with two phases (wet benzene and water with trace benzene).
  • Point b: after adding ethanol, still two phases; ethanol distributes between the two partially-miscible solvents, with greater mole fraction in the water-rich phase.
  • As more ethanol is added, the water-rich phase increases and the benzene-rich phase decreases.
  • Point c: the benzene-rich phase completely disappears, leaving a single liquid phase.
  • The added ethanol has increased the mutual solubilities of benzene and water.

🧂 Two-solid system: NaCl–KCl–water

🧂 Phase regions

  • One-phase area (sln): solution containing all three components.
  • Two-phase areas: solution saturated with respect to one solid salt (either NaCl or KCl) plus that solid.
  • Three-phase area: a triangular region where three phases coexist.

🔺 Eutonic point

The eutonic point is the upper vertex of the three-phase triangle; it represents the composition of solution saturated with respect to both salts.

  • At fixed temperature and pressure, a system of three components and three phases has two degrees of freedom.
  • Since T and p are already fixed, each phase must have a fixed composition.
  • The three vertices of the inner triangle give the fixed compositions: solid NaCl, solid KCl, and solution with x_NaCl = 0.20 and x_KCl = 0.11.

🔗 Common ion effect

  • From the curved boundary separating the one-phase solution area from the two-phase area for solution and solid KCl:
    • Adding NaCl to a saturated solution of KCl decreases the mole fraction of KCl in the saturated solution.
    • Similarly, adding KCl to a saturated solution of NaCl decreases the mole fraction of NaCl.
  • These decreases in solubility when a common ion is added are examples of the common ion effect.
  • Don't confuse: this is not about total amount of dissolved salt, but about the mole fraction of each salt in the saturated solution.

📊 Tie lines in two-phase areas

  • Representative tie lines are drawn in the two-phase areas.
  • One end of a tie line is on the boundary of the one-phase solution area (the saturated solution composition).
  • The other end is at the vertex representing the pure solid salt.
  • The lever rule can be used to find the relative amounts of solution and solid.
88

Cell Diagrams and Cell Reactions

14.1 Cell Diagrams and Cell Reactions

🧭 Overview

🧠 One-sentence thesis

Cell diagrams provide a standardized notation for describing the physical arrangement and chemical reactions of galvanic cells, where the convention of electron flow direction determines how electrode reactions combine into the overall cell reaction.

📌 Key points (3–5)

  • What a galvanic cell is: an electrochemical system with separated reactants that generates electric potential and requires current flow for the reaction to advance.
  • How cell diagrams work: vertical bars separate phases, left/right positioning matters, and the diagram shows both electrodes and their reactants/products.
  • Electrode reaction convention: electrons enter at the right terminal (reduction) and leave at the left terminal (oxidation); the cell reaction is the sum of both electrode reactions with electrons canceling out.
  • Common confusion: cell equilibrium vs reaction equilibrium—a galvanic cell at rest reaches electrochemical equilibrium but not necessarily reaction equilibrium (the reaction would still be spontaneous if reactants were mixed directly).
  • Charge number z: the amount of electrons transferred per unit advancement of the cell reaction, linking chemical change to electrical charge flow.

⚡ What makes a galvanic cell different

⚡ Physical separation of reactants

  • In a normal reaction vessel, reactants and products are in the same phase or in contact, and the reaction proceeds directly until equilibrium.
  • In a galvanic cell, reactants are physically separated so the reaction can only advance when electric current flows through the cell.
  • When the circuit is open and the cell is isolated, the system reaches thermal, mechanical, and transfer equilibrium quickly.

🔋 Cell equilibrium vs reaction equilibrium

Cell equilibrium or electrochemical equilibrium: the state reached when a galvanic cell is isolated with an open circuit.

  • At cell equilibrium, reaction equilibrium is not necessarily present.
  • If you moved the reactants and products to a reaction vessel at the same activities, the reaction might still advance spontaneously.
  • Don't confuse: the cell being "at rest" doesn't mean the reaction has reached its natural equilibrium—it's constrained from proceeding by the lack of current.

📏 What galvanic cells can measure

The excerpt states that measurements of cell potential yield:

  • Precise values of molar reaction quantities
  • Thermodynamic equilibrium constants
  • Mean ionic activity coefficients in electrolyte solutions

🏗️ Physical structure of galvanic cells

🏗️ Components from terminal to terminal

The general arrangement follows this pattern: terminal – electron conductor – ionic conductor(s) – electron conductor – terminal

ComponentDescriptionRole
TerminalsTwo metal wires (usually copper)Pass through system boundary; must be same metal for potential measurement
Electron conductorMetal, graphite, or semiconductorConducts electrons; part of each electrode
Ionic conductorUsually electrolyte solutionAllows ions (not electrons) to move
Liquid junctionContact point between separate ionic conductorsPresent only in cells with transference

🔌 What an electrode is

Electrode or half-cell: the combination of an electron conductor and the ionic conductor in contact with it.

  • Each cell has a left electrode and a right electrode.
  • This left–right distinction establishes an association with reactants and products of the electrode reactions.

🧪 Two types of cells

  • Cell without liquid junction (or without transference): single electrolyte phase with essentially the same composition at both electrodes.
    • Example from excerpt: the hydrogen/silver-silver chloride cell with HCl(aq) as the single ionic conductor.
  • Cell with transference: two separate electrolyte phases separated by a liquid junction.
    • Example from excerpt: zinc–copper (Daniell) cell with separate Zn²⁺(aq) and Cu²⁺(aq) solutions.

📝 Reading cell diagrams

📝 Notation conventions

A cell diagram uses specific symbols:

  • Single vertical bar ( | ): represents a phase boundary
  • Commas: separate different species in the same phase
  • Dashed vertical bar ( ┊ ): represents a liquid junction
  • Pair of dashed vertical bars ( ┊┊ ): liquid junction with negligible potential

📝 Example diagram breakdown

For the cell diagram: Pt | H₂(g) | H⁺(aq), Cl⁻(aq) | AgCl(s) | Ag

Reading left to right:

  • Left electrode: platinum in contact with hydrogen gas and the electrolyte
  • Ionic conductor: aqueous HCl (shown as ions H⁺ and Cl⁻)
  • Right electrode: silver coated with solid AgCl, in contact with the electrolyte
  • The terminals (Cu) may be omitted because cell potential is independent of terminal metal

📝 Why terminals can be omitted

The excerpt notes that copper terminals don't need to appear in the diagram because "the property whose value we seek, the zero-current cell potential, is the same regardless of the metal used for the terminals."

🔄 Writing electrode and cell reactions

🔄 The electron flow convention

All reaction equations follow this convention:

  • Electrons enter at the right terminal → reduction occurs at the right electrode
  • Electrons leave at the left terminal → oxidation occurs at the left electrode

🔄 Steps to write reactions

  1. Left electrode reaction (oxidation): write with electrons as a product (e⁻ on the right side)
  2. Right electrode reaction (reduction): write with electrons as a reactant (e⁻ on the left side)
  3. Balance electron numbers: ensure both reactions have the same absolute value of electron stoichiometric number
  4. Cell reaction: add the two electrode reactions; electrons cancel out

🔄 Example from the hydrogen/silver chloride cell

  • Oxidation at left: H₂(g) → 2H⁺(aq) + 2e⁻
  • Reduction at right: 2AgCl(s) + 2e⁻ → 2Ag(s) + 2Cl⁻(aq)
  • Cell reaction (sum): H₂(g) + 2AgCl(s) → 2H⁺(aq) + 2Cl⁻(aq) + 2Ag(s)

Notice: both electrode reactions use 2 electrons, so they cancel when added.

⚛️ Charge number and advancement

⚛️ Definition of electron number z

Electron number or charge number (z): the amount of electrons entering at the right terminal per unit advancement of the cell reaction.

  • z is a positive dimensionless quantity
  • z equals the absolute value of the stoichiometric number of electrons (|νₑ|) in either electrode reaction
  • Because both electrode reactions are written with the same |νₑ|, all three reactions (left, right, and cell) share the same advancement variable ξ

⚛️ Linking charge to advancement

For an infinitesimal change dξ:

  • Amount of electrons entering at right terminal = z dξ
  • Amount of electrons leaving at left terminal = z dξ (equal amounts)
  • No charge buildup occurs in internal phases

⚛️ The Faraday constant

Faraday constant (F): the charge per amount of protons, equal to the product of elementary charge and Avogadro constant.

  • Value: F = 96,485 C/mol (to five significant figures)
  • Charge per amount of electrons = -F (negative because electrons are negatively charged)
  • The charge entering the right terminal during advancement dξ is: δQ_sys = zF dξ

This equation connects the chemical advancement (ξ) to the electrical charge flow (Q).

89

Electric Potentials in the Cell

14.2 Electric Potentials in the Cell

🧭 Overview

🧠 One-sentence thesis

The equilibrium cell potential of a galvanic cell arises from the algebraic sum of interfacial potential differences at phase boundaries within the cell, which exist because different conducting phases must have different electric potentials to satisfy equilibrium conditions for electron and ion transfer.

📌 Key points (3–5)

  • Cell potential definition: the electric potential difference between terminals of the same metal (right terminal minus left terminal).
  • Where potential differences exist: only at phase boundaries (interfaces) when no current flows; bulk conducting phases have uniform potential.
  • Three types of interfaces: metal–metal contacts, metal–electrolyte interfaces (where electrode reactions occur), and liquid junctions between different electrolyte solutions.
  • Common confusion: the cell reaction is written according to the cell diagram convention (reduction at right, oxidation at left), not according to the direction of spontaneous change—reversing the diagram reverses the sign of the cell potential.
  • Why terminals must be the same metal: if electrons could freely pass through the entire system at the same temperature and composition, both terminals would have the same electron chemical potential and no potential difference would exist.

⚡ Cell potential fundamentals

⚡ Definition and measurement

Cell potential (E_cell): the electric potential difference between terminals of the same metal, defined as φ_R − φ_L (right terminal minus left terminal).

Equilibrium cell potential (E_cell,eq): the cell potential measured under zero-current conditions when the cell is assumed to be in an equilibrium state.

  • The subscripts R and L refer to right and left terminals.
  • Terminals must be the same metal to avoid unknown metal–metal contact potentials in the external measuring circuit.
  • The Galvani potential (inner electric potential) at a point in a phase is interpreted as the average value in a small volume element large enough to contain many molecules.

📏 Measuring equilibrium cell potential

Two practical methods exist:

Potentiometer method:

  • Connect the cell's negative terminal to a battery's negative terminal.
  • Use a slidewire resistor to create a linear potential gradient.
  • Find the position along the slidewire where connecting the cell's positive terminal produces zero current (detected by galvanometer).
  • At this zero-current position, the potential difference equals E_cell,eq.

Modern method:

  • Use a high-impedance digital voltmeter that draws negligible current.
  • This is more convenient than the potentiometer circuit.

🔄 Cell diagram conventions

The cell diagram determines how we write the cell reaction, independent of what actually happens spontaneously:

Example: Zinc–copper cell

  • Diagram written as: Zn | Zn²⁺(aq) | Cu²⁺(aq) | Cu

    • E_cell and E_cell,eq are positive
    • Cell reaction: Zn + Cu²⁺(aq) → Zn²⁺(aq) + Cu (spontaneous forward)
    • Electrons flow left to right through external circuit
  • Diagram reversed as: Cu | Cu²⁺(aq) | Zn²⁺(aq) | Zn

    • E_cell and E_cell,eq are negative
    • Cell reaction: Cu + Zn²⁺(aq) → Cu²⁺(aq) + Zn (written this way even though not spontaneous)
    • Electrons flow right to left through external circuit

Don't confuse: The cell reaction is written according to the cell diagram (reduction at right electrode, oxidation at left electrode), not according to the direction of spontaneous change.

🔌 Origin of cell potential

🔌 Where potential differences exist

When no current flows through the cell:

  • Electric potential is uniform within each bulk conducting phase (otherwise charged particles would spontaneously move).
  • Potential differences exist only at phase boundaries (interfaces).
  • The equilibrium cell potential is the algebraic sum of all interfacial potential differences within the cell.

The potential profile shows vertical steps at each interface, as depicted schematically in the excerpt's Figure 14.4(a).

⚙️ Effect of current flow

When current passes through the cell (external resistor connected):

  • Interfacial potential differences are still present.
  • Internal resistance of electrical conductors causes E_cell to be reduced in magnitude compared to E_cell,eq.
  • The potential profile changes but interfaces still show potential steps.

Important note: Individual interfacial potential differences are theoretical concepts whose values cannot be measured experimentally—only the total cell potential can be measured.

🔗 Metal–metal contact potentials

🔗 What causes contact potentials

Contact potential: an electric potential difference at an interface between two different metals.

When two different metals are placed in contact:

  • Local densities of free (mobile) electrons change.
  • An electrical double layer forms with excess positive charge on one side and excess negative charge on the other side.
  • This double layer creates the contact potential.

⚖️ Equilibrium condition for electrons

The chemical potential of free electrons in a metal phase depends on the electric potential of the phase:

Relationship: μ_e(φ) = μ_e(0) − F·φ

Where:

  • μ_e(φ) is the electron chemical potential at electric potential φ
  • μ_e(0) is the electron chemical potential at zero electric potential (depends only on temperature and composition)
  • F is the Faraday constant (96,485 C/mol)

Electron transfer equilibrium between two phases requires equal electron chemical potentials:

Result: φ″ − φ′ = [μ_e″(0) − μ_e′(0)] / F

This shows the contact potential depends only on temperature and the compositions of the two metals.

🚫 Why terminals must be the same metal

If electrons were free to pass from one terminal through the system to the other terminal of the same temperature and composition:

  • In a zero-current equilibrium state, μ_e would be the same in both terminals.
  • There would be no potential difference between the terminals.
  • The system would not function as a galvanic cell.

Conclusion: A galvanic cell must have at least one electrical conductor that is not an electron conductor (i.e., an ionic conductor like an electrolyte solution).

🧪 Metal–electrolyte interfaces

🧪 Electrode reaction equilibrium

An electrode reaction takes place at the interface between a metal electron conductor and an electrolyte solution.

Equilibrium condition: The sum of (stoichiometric coefficient × chemical potential) over all reactants and products equals zero: Σ ν_i μ_i = 0

Since chemical potentials of ions and electrons depend on electric potentials of their phases, the metal and solution must generally have different electric potentials for this sum to be zero.

🔋 Example: Copper electrode

For the copper electrode reaction: Cu²⁺(aq) + 2e⁻(Cu) → Cu

Equilibrium requires: μ(Cu) − μ(Cu²⁺) − 2μ_e(Cu) = 0

The interfacial potential difference between the copper conductor and the solution must satisfy this condition.

🌊 Structure of the interface

The interfacial potential difference arises from:

  • Charge separation across the interface
  • Orientation of polar molecules on the solution side
  • Specific adsorption of ions

Thickness of zones where properties differ from bulk phases:

  • Metal side: no greater than 10⁻¹¹ m
  • Solution side: no greater than 10⁻⁷ m

💧 Liquid junction potentials

💧 What causes liquid junction potentials

Some galvanic cells contain two electrolyte solutions with different compositions, separated by a porous barrier or junction to prevent rapid mixing.

Liquid junction potential: an electric potential difference at the junction between two electrolyte solutions, caused by diffusion of ions between the two bulk phases.

🌀 How it arises

Imagine both solution phases initially had the same electric potential:

  1. An ion species with different chemical potentials in the two solutions would spontaneously diffuse toward lower chemical potential.
  2. Different ions diffuse at different rates.
  3. This results in net charge transfer across the junction.
  4. An electric potential difference develops.
  5. In the equilibrium state, this potential difference prevents further net charge transfer under zero-current conditions.

🧂 Salt bridges

A liquid junction may consist of a bridging solution in a salt bridge.

Common type:

  • Glass tube filled with gel made from agar and concentrated aqueous KCl or KNO₃
  • Believed to reduce the liquid junction potential to several millivolts or less

Cell diagram notation:

  • Liquid junction with non-negligible potential: single dashed vertical bar |
  • Liquid junction with negligible potential: pair of dashed vertical bars ||

Example: Zn | Zn²⁺(aq) || Cu²⁺(aq) | Cu (negligible junction potential assumed)

⚗️ Advancement and charge

⚗️ Electron number definition

Electron number (charge number) z: the amount of electrons entering at the right terminal per unit advancement of the cell reaction.

  • z is a positive dimensionless quantity.
  • Equal to the absolute value of the stoichiometric number of electrons in either electrode reaction.
  • Both electrode reactions are written with the same value of |ν_e|, so their advancements and the cell reaction advancement are all described by the same advancement variable ξ.

🔋 Faraday constant

Faraday constant F: a physical constant defined as the charge per amount of protons, equal to the product of the elementary charge (charge of a proton) and the Avogadro constant: F = e·N_A.

Value: F = 96,485 C/mol (to five significant figures)

The charge per amount of electrons is −F.

⚡ Charge transfer during reaction

For an infinitesimal advancement dξ:

  • An amount of electrons equal to z·dξ enters the system at the right terminal.
  • An equal amount of electrons leaves at the left terminal.
  • There is no buildup of charge in any internal phases.

Charge entering the right terminal: δQ_sys = z·F·dξ

This relationship connects the chemical advancement of the cell reaction to the electrical charge transferred.

90

Molar Reaction Quantities of the Cell Reaction

14.3 Molar Reaction Quantities of the Cell Reaction

🧭 Overview

🧠 One-sentence thesis

The molar reaction Gibbs energy of a cell reaction is directly related to the equilibrium cell potential, allowing precise measurement of thermodynamic quantities like standard Gibbs energies and equilibrium constants through electrochemical methods.

📌 Key points (3–5)

  • Core relationship: The molar reaction Gibbs energy of the cell reaction (ΔᵣGcell) equals −zFEcell,eq, where z is the number of electrons transferred, F is Faraday's constant, and Ecell,eq is the equilibrium cell potential.
  • Distinction between cell and direct reactions: ΔᵣGcell (cell reaction) may differ from ΔᵣG (direct reaction) in cells with liquid junctions due to different electric potentials affecting ionic chemical potentials.
  • Liquid junction effects: Cells without liquid junctions have ΔᵣGcell = ΔᵣG; cells with liquid junctions have ΔᵣGcell = ΔᵣG − zFEⱼ, where Eⱼ is the liquid junction potential.
  • Common confusion: The cell reaches full equilibrium (including reaction equilibrium) only when it is "dead" (exhausted); zero current alone means thermal, mechanical, and transfer equilibrium but not necessarily reaction equilibrium.
  • Practical application: Standard cell potentials allow noncalorimetric determination of standard Gibbs energies and equilibrium constants through precise voltage measurements.

⚡ Fundamental relationship between Gibbs energy and cell potential

⚡ Defining the molar reaction Gibbs energy of the cell reaction

ΔᵣGcell: the molar reaction Gibbs energy of a cell reaction, defined by the sum over all reactants and products of (stoichiometric number × chemical potential).

  • Written as: ΔᵣGcell = Σᵢ νᵢμᵢ, where the sum covers all species in the cell reaction.
  • Also equals the partial derivative (∂Gcell/∂ξ) at constant temperature and pressure, where ξ is the advancement of the cell reaction.
  • Key distinction: This is different from ΔᵣG of the direct reaction because ionic chemical potentials depend on the electric potential of their phase.

🔋 Deriving the core equation

When a galvanic cell is in zero-current equilibrium:

  • Both electrode reactions are at equilibrium.
  • At the left electrode: electrons are products with stoichiometric number z.
  • At the right electrode: electrons are reactants with stoichiometric number −z.

The equilibrium conditions are:

  • Left electrode: Σᵢ νᵢμᵢ + zμₑ(LE) = 0
  • Right electrode: Σⱼ νⱼμⱼ − zμₑ(RE) = 0

Adding these equations and recognizing that the chemical species sums equal ΔᵣGcell:

  • ΔᵣGcell = z[μₑ(LE) − μₑ(RE)]

Since electron transfer equilibrium exists between conductors and terminals, and using the relationship between chemical potential and electric potential:

  • ΔᵣGcell = −zFEcell,eq

This is the fundamental equation connecting thermodynamics and electrochemistry.

🔌 Independence from terminal composition

Important deductions for cells with identical terminal metals:

  • The equilibrium cell potential Ecell,eq does not depend on which metal is used for the terminals.
  • Interposing any metal conductor between an electrode and terminal does not affect Ecell,eq.
  • The value of ΔᵣGcell has nothing to do with terminal composition (from the definition as a sum over chemical potentials).

Alternative derivation: Using reversible electrical work at constant T and p equals Gibbs energy change (δwₑₗ,ᵣₑᵥ = dGcell), substituting δwₑₗ,ᵣₑᵥ = Ecell,eq δQsys and δQsys = −zF dξ, then dividing by dξ gives the same result.

Note: This derivation strictly applies only to cells without liquid junctions, as cells with liquid junctions are not truly reversible (different ions carry current in different directions).

🔄 Comparing cell reactions and direct reactions

🔄 The thought experiment setup

To understand the relationship between ΔᵣGcell and ΔᵣG:

  1. Start with a galvanic cell in zero-current equilibrium with defined temperature, pressure, and composition.
  2. Imagine a separate reaction vessel at the same T and p, containing the same reactants and products at the same activities.
  3. In the reaction vessel, species are in direct contact (no electrical circuit), so spontaneous direct reaction can occur.

Example: For a zinc-copper cell, the reaction vessel would have zinc and copper strips in contact with a solution containing both ZnSO₄ and CuSO₄.

The reaction equations use the same stoichiometric numbers νᵢ in both systems, but chemical potentials of charged species may differ due to different electric potentials.

🧪 Cells without liquid junctions

When all ions are in a single electrolyte solution phase:

  • The same solution phase exists in both the cell and the reaction vessel.
  • When all ions are in the same phase, the sum Σᵢ νᵢμᵢ is independent of the electric potentials of any phases.
  • Therefore: ΔᵣGcell = ΔᵣG (no liquid junction)

This equality allows direct comparison between electrochemical and thermodynamic quantities.

🌉 Cells with liquid junctions

For cells with two electrolyte solutions separated by a liquid junction:

  • Write ΔᵣGcell = Σᵢ νᵢμᵢ(i) + Σⱼ νⱼμⱼ(j), where i indexes species at the left electrode and j at the right.
  • Substitute μᵢ(φ) = μᵢ(0) + zᵢFφ for each ionic species.
  • The sums Σᵢ νᵢμᵢ(0) + Σⱼ νⱼμⱼ(0) equal ΔᵣG (molar reaction Gibbs energy in phases of zero potential).
  • Using charge conservation (Σᵢ νᵢzᵢ − z = 0 and Σⱼ νⱼzⱼ + z = 0):
  • ΔᵣGcell = ΔᵣG − zFEⱼ (cell with liquid junction)

where Eⱼ = φ″ − φ′ is the liquid junction potential.

📐 Practical formulas for equilibrium cell potential

Combining with ΔᵣGcell = −zFEcell,eq:

Cell typeEquationApplication
Without liquid junctionEcell,eq = −ΔᵣG/(zF)Precise evaluation of ΔᵣG from measured Ecell,eq
With liquid junctionEcell,eq = −ΔᵣG/(zF) + EⱼRequires negligible or estimated Eⱼ

Don't confuse: Zero current means thermal, mechanical, and transfer equilibrium, but the cell has full reaction equilibrium only when ΔᵣG = 0, which requires Ecell,eq = 0 (no liquid junction) or Ecell,eq = Eⱼ (with liquid junction)—this is a "dead" cell that can no longer do electrical work.

🎯 Standard quantities and equilibrium constants

🎯 Defining the standard cell potential

Standard cell potential (E°cell,eq): the equilibrium cell potential of a hypothetical cell in which each reactant and product is in its standard state at unit activity, with negligible liquid junction potential if a junction is present.

  • This is a hypothetical reference state, not necessarily an achievable experimental condition.
  • In this state, ΔᵣGcell equals the standard molar reaction Gibbs energy ΔᵣG°.
  • From the fundamental equation: ΔᵣG° = −zFE°cell,eq

🔬 Noncalorimetric determination of equilibrium constants

Since ΔᵣG° = −RT ln K (where K is the thermodynamic equilibrium constant):

  • ln K = (zF/RT)E°cell,eq

This equation allows evaluation of K through precise electrochemical measurements rather than calorimetry.

Example: For the cell Ag | Ag⁺(aq) | Cl⁻(aq) | AgCl(s) | Ag (where || indicates negligible liquid junction):

  • Electrode reactions:
    • Ag(s) → Ag⁺(aq) + e⁻
    • AgCl(s) + e⁻ → Ag(s) + Cl⁻(aq)
  • Cell reaction: AgCl(s) → Ag⁺(aq) + Cl⁻(aq)
  • The equilibrium constant is the solubility product Kₛ of silver chloride.
  • At 298.15 K, E°cell,eq = −0.5770 V.
  • Use ln K = (zF/RT)E°cell,eq to calculate Kₛ.

📊 Advantages of the electrochemical method

  • High precision: Cell potentials can be measured with great accuracy.
  • Direct thermodynamic information: Provides ΔᵣG° and K without calorimetric measurements.
  • Requirement: The reaction must be carried out in a galvanic cell, preferably without a liquid junction or with negligible Eⱼ.

Limitation: For cells with liquid junctions, the liquid junction potential Eⱼ must be negligible or reliably estimated from theory (e.g., using a salt bridge with concentrated KCl or KNO₃, which reduces Eⱼ to several millivolts or less).

91

The Nernst Equation

14.4 The Nernst Equation

🧭 Overview

🧠 One-sentence thesis

The Nernst equation relates the equilibrium cell potential of a galvanic cell to the standard cell potential and the activities of reactants and products, showing how cell potential changes with composition and predicting when a cell becomes "dead" at reaction equilibrium.

📌 Key points (3–5)

  • What the Nernst equation does: connects the measured equilibrium cell potential E_cell,eq to the standard cell potential E°_cell,eq and the reaction quotient Q_rxn (which depends on activities of reactants and products).
  • Standard cell potential: the equilibrium potential when all reactants and products are in their standard states (unit activity) with negligible liquid junction potential; it depends only on temperature.
  • How activities affect potential: decreasing product activities or increasing reactant activities increases E_cell,eq (greater tendency for spontaneous forward reaction); at reaction equilibrium Q_rxn equals K and E_cell,eq becomes zero (dead cell).
  • Common confusion: E°_cell,eq is a constant at a given temperature, but E_cell,eq varies with the actual activities in the cell.
  • Practical use: the Nernst equation allows calculation of activity coefficients from measured cell potentials, and evaluation of equilibrium constants from standard cell potentials.

⚡ Standard cell potential and its meaning

⚡ Definition of standard cell potential

Standard cell potential E°_cell,eq: the equilibrium cell potential of a hypothetical galvanic cell in which each reactant and product of the cell reaction is in its standard state at unit activity, and any liquid junction has negligible liquid junction potential.

  • E°_cell,eq is a function only of temperature for a given cell reaction with given choices of standard states.
  • It does not depend on the actual activities of species in a real cell.
  • The excerpt emphasizes this is a hypothetical reference state.

🔗 Relationship to thermodynamic quantities

The standard cell potential connects directly to standard molar reaction Gibbs energy:

Δ_r G° = –zFE°_cell,eq

where z is the number of electrons transferred and F is Faraday's constant.

This relationship allows evaluation of:

  • Equilibrium constant K: ln K = (zF/RT) E°_cell,eq
  • Standard molar reaction enthalpy: Δ_r H° = zF[T(dE°_cell,eq/dT) – E°_cell,eq]
  • Standard molar reaction entropy: Δ_r S° = zF(dE°_cell,eq/dT)

Example: For the cell reaction AgCl(s) → Ag⁺(aq) + Cl⁻(aq), the equilibrium constant is the solubility product K_s of silver chloride. At 298.15 K, E°_cell,eq = –0.5770 V allows calculation of K_s.

🔍 Standard vs actual cell potential

Don't confuse:

  • E°_cell,eq: constant at a given temperature, hypothetical standard-state condition
  • E_cell,eq: measured equilibrium cell potential of an actual cell, depends on activities, temperature, and liquid junction potential (if present)

📐 The Nernst equation formula

📐 General form

The Nernst equation is:

E_cell,eq = E°_cell,eq – (RT/zF) ln Q_rxn

(valid for cells without liquid junction, or with negligible liquid junction potential E_j = 0)

Where:

  • Q_rxn is the reaction quotient: the product of activities of products (each raised to its stoichiometric coefficient) divided by the product of activities of reactants (each raised to its stoichiometric coefficient)
  • z is the number of electrons transferred in the cell reaction
  • R is the gas constant, T is temperature, F is Faraday's constant

🌡️ Simplified form at 298.15 K

At T = 298.15 K (25.00 °C), RT/F = 0.02569 V, so:

E_cell,eq = E°_cell,eq – (0.02569 V / z) ln Q_rxn

This compact form is convenient for calculations at room temperature.

🔄 How activities affect cell potential

🔄 Effect of changing activities

The Nernst equation shows how composition changes affect the cell potential:

Change in activitiesEffect on ln Q_rxnEffect on E_cell,eqPhysical meaning
Decrease product activitiesDecreases ln Q_rxnIncreases E_cell,eqGreater tendency for forward reaction
Increase reactant activitiesDecreases ln Q_rxnIncreases E_cell,eqGreater tendency for forward reaction
Increase product activitiesIncreases ln Q_rxnDecreases E_cell,eqLess tendency for forward reaction
Decrease reactant activitiesIncreases ln Q_rxnDecreases E_cell,eqLess tendency for forward reaction

The excerpt explains: "E_cell,eq should be greater when the forward cell reaction has a greater tendency for spontaneity."

⚖️ Special case: standard state

When each reactant and product is in its standard state:

  • Each activity equals unity (1)
  • ln Q_rxn = 0 (since ln 1 = 0)
  • Therefore E_cell,eq = E°_cell,eq

This confirms that the standard cell potential is indeed the potential under standard-state conditions.

⚰️ Special case: reaction equilibrium (dead cell)

When the cell reaction reaches equilibrium (e.g., if terminals are short-circuited):

  • Q_rxn becomes equal to the thermodynamic equilibrium constant K
  • The Nernst equation becomes: E_cell,eq = E°_cell,eq – (RT/zF) ln K
  • But (RT/zF) ln K equals E°_cell,eq (from the relationship ln K = (zF/RT) E°_cell,eq)
  • Therefore E_cell,eq = 0

The excerpt states: "the cell is 'dead' and is incapable of performing electrical work on the surroundings."

Don't confuse: A dead cell (E_cell,eq = 0) is at reaction equilibrium, not necessarily at standard state.

🧪 Application example: HCl cell

🧪 The cell reaction

The excerpt illustrates the Nernst equation with the reaction:

H₂(g) + 2 AgCl(s) → 2 H⁺(aq) + 2 Cl⁻(aq) + 2 Ag(s)

This reaction takes place in a cell without liquid junction, with aqueous HCl as the electrolyte solution.

🧪 Constructing the reaction quotient

The reaction quotient is:

Q_rxn = (a²_H⁺ · a²_Cl⁻ · a²_Ag) / (a_H₂ · a²_AgCl)

Simplifications (with negligible error):

  • Activities of solids (Ag, AgCl) are approximately 1
  • Solute activities: a_H⁺ = γ_+ m_+/m°, a_Cl⁻ = γ_– m_–/m°
  • Hydrogen activity: a_H₂ = f_H₂/p° (fugacity over standard pressure)
  • Ion molalities m_+ and m_– both equal the HCl molality m_B

This gives:

Q_rxn = (γ⁴_± · (m_B/m°)⁴) / (f_H₂/p°)

where γ_± is the mean ionic activity coefficient of HCl.

🧪 The Nernst equation for this cell

Substituting into the Nernst equation (with z = 2 electrons):

E_cell,eq = E°cell,eq – (RT/2F) ln[(γ⁴± · (m_B/m°)⁴) / (f_H₂/p°)]

Expanding the logarithm:

E_cell,eq = E°cell,eq – (2RT/F) ln γ± – (2RT/F) ln(m_B/m°) + (RT/2F) ln(f_H₂/p°)

🧪 Determining activity coefficients

By measuring E_cell,eq for a cell with known values of m_B and f_H₂, and with a derived value of E°cell,eq, this equation allows calculation of the mean ionic activity coefficient γ± of the HCl solute.

The excerpt notes: "This is how the experimental curve for aqueous HCl in Fig. 10.3 on page 298 was obtained."

This demonstrates a practical noncalorimetric method for determining activity coefficients from electrochemical measurements.

92

Evaluation of the Standard Cell Potential

14.5 Evaluation of the Standard Cell Potential

🧭 Overview

🧠 One-sentence thesis

The standard cell potential can be determined experimentally by measuring cell potentials at various electrolyte concentrations and extrapolating to infinite dilution where ionic activity coefficients become unity.

📌 Key points

  • What the standard cell potential is: a temperature-dependent thermodynamic quantity that can be evaluated by extrapolation to infinite dilution.
  • Why extrapolation is needed: at infinite dilution, ionic activity coefficients approach unity, simplifying the calculation.
  • How the method works: measure cell potential at different molalities, apply the Debye–Hückel formula, and extrapolate a modified quantity (E′ cell) to zero molality.
  • Common confusion: the standard cell potential itself is independent of how we write the reaction equation (same physical cell), but molar reaction quantities like delta-r-G do depend on stoichiometric coefficients.
  • Practical outcome: this extrapolation method yields highly precise values (uncertainty around 0.1 mV) for standard cell potentials.

🔬 The extrapolation method

🧪 Starting equation

The Nernst equation for the example cell reaction (H₂ + 2AgCl → 2H⁺ + 2Cl⁻ + 2Ag) can be rearranged to isolate the standard cell potential:

E°(cell,eq) = E(cell,eq) + (2RT/F) ln γ(±) + (2RT/F) ln (m_B/m°) - (RT/2F) ln (f_H₂/p°)

  • E(cell,eq) is the measured equilibrium cell potential
  • γ(±) is the mean ionic activity coefficient of the HCl electrolyte
  • m_B is the molality of HCl
  • f_H₂ is the fugacity of hydrogen gas

🎯 The challenge

  • All quantities on the right side can be measured except the mean ionic activity coefficient γ(±)
  • We cannot know the exact value of ln γ(±) at any given molality until we already know E°(cell,eq)
  • This creates a circular dependency problem

🔑 The solution: Debye–Hückel approximation

The method uses the Debye–Hückel formula to approximate ln γ(±):

  • The Debye–Hückel formula becomes more accurate as ionic strength (or molality) decreases
  • As m_B approaches zero, γ(±) approaches unity and ln γ(±) approaches zero
  • This behavior allows extrapolation to infinite dilution

📐 Defining the extrapolation function

📊 The modified quantity E′ cell

A new quantity E′ cell is defined:

E′ cell = E(cell,eq) + (2RT/F) × [Debye–Hückel term] + (2RT/F) ln (m_B/m°) - (RT/2F) ln (f_H₂/p°)

Where the Debye–Hückel term is: A√(m_B) / (1 + Ba√(m_B))

  • A and B are known constants at any temperature
  • a is an ion-size parameter (a reasonable value can be chosen)

🔄 Why this works

  • The difference between the right side of the original equation (14.5.1) and the E′ cell equation (14.5.2) consists of contributions to ln γ(±) not accounted for by the Debye–Hückel formula
  • These unaccounted contributions approach zero as molality approaches zero (infinite dilution)
  • Therefore, extrapolating measured E′ cell values to m_B = 0 yields the true E°(cell,eq)

📈 Practical implementation

  • Measure E(cell,eq) at various HCl molalities
  • Calculate E′ cell for each molality using equation 14.5.2
  • Plot E′ cell versus m_B
  • Extrapolate to m_B = 0 (often using a linear fit)
  • The intercept gives E°(cell,eq)

Example: Figure 14.5 shows this extrapolation for the H₂/AgCl cell at 298.15 K, yielding E°(cell,eq) = 0.2222 V with uncertainty around 0.1 mV.

⚖️ Effect of stoichiometric coefficients

🔢 Multiplying the reaction equation

The excerpt addresses what happens when all stoichiometric coefficients are multiplied by the same constant.

Original reaction:

  • H₂(g) + 2AgCl(s) → 2H⁺(aq) + 2Cl⁻(aq) + 2Ag(s)
  • Number of electrons transferred: z = 2

Halved reaction:

  • ½H₂(g) + AgCl(s) → H⁺(aq) + Cl⁻(aq) + Ag(s)
  • Number of electrons transferred: z = 1

🎭 What changes and what doesn't

QuantityDoes it change?Why
E(cell,eq)NoThe physical cell is the same; measurable quantities are unaffected
E°(cell,eq)NoSame reason—it's a property of the physical system
Nernst equation formYesThe value of z changes, altering the equation structure
Molar reaction quantities (Δ_r G, Δ_r G°)YesThese are defined per extent of reaction, which depends on stoichiometry

⚠️ Don't confuse

  • The standard cell potential is independent of how we write the reaction equation
  • Molar Gibbs energy changes do depend on stoichiometric coefficients because they represent changes per mole of reaction progress

🌡️ Temperature dependence

📏 Key characteristic

The standard cell potential E°(cell,eq) for a given cell reaction depends only on temperature.

  • It does not depend on concentration, pressure (beyond standard state definition), or how the reaction equation is written
  • This makes it a fundamental thermodynamic property at each temperature
  • The excerpt emphasizes this is why E°(cell,eq) has "useful thermodynamic applications"
93

Standard Electrode Potentials

14.6 Standard Electrode Potentials

🧭 Overview

🧠 One-sentence thesis

Standard electrode potentials allow us to calculate standard cell potentials for any galvanic cell without measuring each one individually, by using the hydrogen electrode as a universal reference point.

📌 Key points (3–5)

  • What standard electrode potentials are: the standard cell potential of a cell with a hydrogen electrode on the left and the electrode of interest on the right.
  • The hydrogen reference: the standard hydrogen electrode is defined to have zero potential at all temperatures, serving as the universal reference.
  • Key calculation formula: E°(cell,eq) = E°(R) − E°(L), where R is the right electrode and L is the left electrode.
  • Common confusion: the standard hydrogen electrode is hypothetical (cannot actually be constructed) because it requires species in their hypothetical standard states.
  • Why it matters: measuring just 10 electrodes allows calculation of standard cell potentials for 35 additional cells without further experiments.

🔋 The standard hydrogen electrode reference

🔋 What the standard hydrogen electrode is

Standard hydrogen electrode: a hydrogen electrode in which H₂(g) and H⁺(aq) are in their standard states.

  • This electrode cannot actually be built in the laboratory because it requires hypothetical gas and solute standard states.
  • Despite being hypothetical, it serves as the universal reference for all electrode potential measurements.

🎯 Zero potential by convention

  • The standard electrode potential of the hydrogen electrode is defined as zero at all temperatures.
  • This is a convention, not a measured value—it establishes the reference point for all other electrodes.
  • Any cell with hydrogen electrodes at both left and right has a standard cell potential of zero.

📐 Defining standard electrode potentials

📐 How standard electrode potentials are defined

Standard electrode potential E°: the standard cell potential of a cell with a hydrogen electrode at the left and the electrode of interest at the right.

  • The definition always places the hydrogen electrode on the left side of the cell diagram.
  • The electrode of interest is placed on the right side.
  • Example: For the cell Pt | H₂(g) | HCl(aq) | AgCl(s) | Ag, the standard electrode potential of the silver–silver chloride electrode equals the standard cell potential of this entire cell.

🌡️ Temperature dependence

  • Standard electrode potentials are functions of temperature only.
  • They are nonzero for all electrodes except the hydrogen electrode.
  • The excerpt mentions uncertainty on the order of only 0.1 mV for careful measurements.

🧮 Calculating standard cell potentials

🧮 The fundamental formula

The key relationship for any galvanic cell is:

E°(cell,eq) = E°(R) − E°(L)

where:

  • E°(cell,eq) is the standard cell potential of the cell
  • E°(R) is the standard electrode potential of the right electrode
  • E°(L) is the standard electrode potential of the left electrode

🔬 Derivation using three cells

The excerpt derives this formula by considering three cells:

CellLeft electrodeRight electrodeStandard cell potential
Cell 1Electrode LElectrode RE°(cell,eq) (what we want)
Cell 2HydrogenElectrode LE°(L)
Cell 3HydrogenElectrode RE°(R)

How the derivation works:

  • Write cell reactions for cells 1 and 2 using the same electron number z
  • The sum of reactions 1 and 2 equals reaction 3
  • An infinitesimal advancement dξ of reaction 1 combined with an equal advancement of reaction 2 causes the same changes in amounts as advancement dξ of reaction 3
  • Therefore: ΔᵣG°(reaction 1) + ΔᵣG°(reaction 2) = ΔᵣG°(reaction 3)
  • Substituting ΔᵣG° = −zFE°(cell,eq) for each reaction gives: E°(cell,eq) + E°(L) = E°(R)
  • Rearranging yields: E°(cell,eq) = E°(R) − E°(L)

⚠️ Important requirement

Don't confuse: The cell reactions must be written using the same value of the electron number z for the relationship to hold. This ensures the advancements are directly comparable.

💡 Practical advantages

💡 Efficiency of the method

The excerpt emphasizes the enormous experimental savings:

  • Measuring E°(cell,eq) for just 10 different cells (only one needs to include a hydrogen electrode) provides values of E° for 10 electrodes
  • From these 10 electrode potentials, you can calculate E°(cell,eq) for 35 additional cells without hydrogen electrodes
  • This avoids the "involved experimental procedure" of measuring equilibrium cell potential at different electrolyte molalities for each individual cell

🔄 Flexibility in measurement

  • Neither electrode in a measured cell has to be a hydrogen electrode
  • The hydrogen electrode is "difficult to work with experimentally"
  • You can calculate standard electrode potentials from cells containing any two electrodes, as long as you know the standard electrode potential of one of them

🧪 General applicability

  • Equation 14.6.3 is described as "a general relation applicable to any galvanic cell"
  • This universality makes it a powerful tool for electrochemical calculations
  • Example: Once you have a table of standard electrode potentials, you can predict the behavior of countless cell combinations without further experiments