| Non-Rationalised Science NCERT Notes and Solutions (Class 6th to 10th) | ||||||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| 6th | 7th | 8th | 9th | 10th | ||||||||||
| Non-Rationalised Science NCERT Notes and Solutions (Class 11th) | ||||||||||||||
| Physics | Chemistry | Biology | ||||||||||||
| Non-Rationalised Science NCERT Notes and Solutions (Class 12th) | ||||||||||||||
| Physics | Chemistry | Biology | ||||||||||||
Chapter 6 Thermodynamics
The energy stored within the chemical bonds of molecules is a form of chemical energy. This energy can be released as heat during chemical reactions (e.g., combustion of fuels like methane or coal) or converted into other forms of energy, such as mechanical work (in engines) or electrical energy (in batteries).
Thermodynamics is the branch of science that studies these energy transformations and the relationships between different forms of energy. It focuses on macroscopic systems (dealing with large numbers of molecules) and is concerned with the initial and final states of a system undergoing a change, rather than the specific pathway or rate of the change.
The laws of thermodynamics apply to systems that are in equilibrium or moving between equilibrium states. Key questions addressed by thermodynamics include:
- How can we quantify the energy changes that occur during chemical reactions or physical processes?
- Can we predict whether a specific reaction or process will occur spontaneously?
- To what extent will a reaction proceed before reaching equilibrium?
Thermodynamic Terms
To study energy changes in chemical systems, we need to define certain thermodynamic terms.
The System And The Surroundings
- A system is the specific part of the universe that we are observing or studying.
- The surroundings are everything else in the universe outside the system.
- The system and the surroundings together constitute the universe.
The Universe = System + Surroundings
Practically, the surroundings are considered to be the portion of the universe near the system that can interact with it.
The boundary is the real or imaginary surface that separates the system from the surroundings. It allows for tracking the exchange of matter and energy.
Types Of The System
Systems are classified based on whether they exchange matter and/or energy with their surroundings.
- Open System: Exchanges both energy and matter with the surroundings. Example: Reactants in an open beaker.
- Closed System: Exchanges energy but not matter with the surroundings. Example: Reactants in a sealed vessel made of a conducting material (like metal).
- Isolated System: Exchanges neither energy nor matter with the surroundings. Example: Reactants in a sealed, insulated vessel (like a thermos flask).
The State Of The System
The state of a thermodynamic system is defined by its macroscopic (measurable) properties. For a gas, its state can be described by specifying its pressure ($p$), volume ($V$), temperature ($T$), and amount (number of moles, $n$).
These properties (p, V, T, n, etc.) are called state variables or state functions. Their value depends only on the current state of the system, regardless of how that state was reached (i.e., they are path-independent). Once a minimum number of state variables are fixed, the others are automatically determined.
The Internal Energy As A State Function
The total energy of a system, encompassing all forms of energy within it (kinetic energy of particles, potential energy due to intermolecular forces, etc.), is called its Internal Energy ($U$). We cannot measure the absolute value of internal energy, but we can measure the change in internal energy ($\Delta U$) when a system transitions from one state to another.
The internal energy of a system can change through the exchange of work ($w$) or heat ($q$) with the surroundings, or by the transfer of matter (in open systems).
Let's consider an adiabatic process, where no heat exchange occurs between the system and surroundings ($q=0$). This requires an adiabatic wall (a boundary that prevents heat flow).
If we do work on an adiabatic system (e.g., stirring water in a thermos flask), its temperature increases. Experiments by J.P. Joule showed that the same amount of work done on the system, regardless of how it was done, produces the same change in temperature and thus the same change in internal energy. The change in internal energy in an adiabatic process ($\Delta U$) is equal to the adiabatic work ($w_{ad}$):
$\Delta U = w_{ad}$ (for an adiabatic change, $q=0$)
This demonstrates that internal energy ($U$) is a state function, as the change $\Delta U$ depends only on the initial and final states, not the process (path) by which work was done.
Conventionally, work done on the system is positive ($w > 0$), increasing the internal energy ($\Delta U > 0$). Work done by the system is negative ($w < 0$), decreasing internal energy ($\Delta U < 0$).
The internal energy can also change by the transfer of heat ($q$) through thermally conducting walls (instead of adiabatic walls) without work being done ($w=0$).
If heat is absorbed by the system at constant volume ($w=0$), its internal energy increases. The change in internal energy is equal to the heat absorbed ($q_V$):
$\Delta U = q_V$ (for a process at constant volume, $w=0$)
Conventionally, heat transferred to the system is positive ($q > 0$), increasing internal energy. Heat transferred from the system is negative ($q < 0$), decreasing internal energy.
(c) The general case - First Law of Thermodynamics
In the most general case, a change in the system's state can involve both work ($w$) and heat ($q$) exchange with the surroundings. The change in internal energy is the sum of the heat absorbed by the system and the work done on the system:
$\Delta U = q + w$
This is the mathematical statement of the First Law of Thermodynamics. It states that the energy of an isolated system is constant, reflecting the law of conservation of energy: energy cannot be created or destroyed, only transformed from one form to another.
For a given change of state, the individual values of $q$ and $w$ depend on the path taken, but their sum ($q+w$) is always equal to $\Delta U$, which is path-independent because $U$ is a state function.
Problem 6.1. Express the change in internal energy of a system when
(i) No heat is absorbed by the system from the surroundings, but work (w) is done on the system. What type of wall does the system have ?
(ii) No work is done on the system, but q amount of heat is taken out from the system and given to the surroundings. What type of wall does the system have?
(iii) w amount of work is done by the system and q amount of heat is supplied to the system. What type of system would it be?
Answer:
We use the First Law of Thermodynamics: $\Delta U = q + w$.
(i) "No heat is absorbed by the system" means $q=0$. "work (w) is done on the system" means $w > 0$ (by convention). So, $\Delta U = 0 + w = w$. A system with no heat exchange has an adiabatic wall (thermally insulating boundary). $\Delta U = w_{ad}$.
(ii) "No work is done on the system" means $w=0$. "q amount of heat is taken out from the system and given to the surroundings" means the system loses heat, so $q$ for the system is negative. Let the amount of heat taken out be $|q|$. Then $q = -|q|$. So, $\Delta U = -|q| + 0 = -|q|$. Walls allowing heat exchange are thermally conducting walls.
(iii) "w amount of work is done by the system" means $w < 0$. Let the amount of work done by the system be $|w|$. Then $w = -|w|$. "q amount of heat is supplied to the system" means the system gains heat, so $q > 0$. Let the amount of heat supplied be $|q|$. Then $q = |q|$. So, $\Delta U = |q| + (-|w|) = q - w$. A system that exchanges both heat and work (but not matter) is a closed system.
Applications
Understanding how energy changes occur as heat or work allows us to quantify these changes in chemical reactions and processes.
Work
In thermodynamics, we often consider pressure-volume (PV) work, which is work done by or on a system due to a change in its volume against an external pressure. Consider a gas in a cylinder with a frictionless piston.
If the gas expands against a constant external pressure ($p_{ex}$), the system does work on the surroundings. If the gas is compressed by a constant external pressure, the surroundings do work on the system.
Let the initial volume be $V_i$ and the final volume be $V_f$. The change in volume is $\Delta V = V_f - V_i$. If the piston moves a distance $l$ and has area $A$, $\Delta V = A \times l$. The force exerted by the external pressure is $F = p_{ex} \times A$.
Work ($w$) is force times distance. Work done on the system by compression is:
$w = F \times (-l)$ (distance moved is against the direction of force exerted by the system)
$w = (p_{ex} \times A) \times (-l)$
$w = -p_{ex} \times (A \times l) = -p_{ex} \Delta V = -p_{ex}(V_f - V_i)$
The negative sign is included so that $w$ has the correct thermodynamic sign convention: if the gas is compressed, $V_f < V_i$, $\Delta V$ is negative, and $w = -p_{ex}(-\Delta V) = +p_{ex}|\Delta V|$ is positive (work done on the system). If the gas expands, $V_f > V_i$, $\Delta V$ is positive, and $w = -p_{ex}(+\Delta V) = -p_{ex}|\Delta V|$ is negative (work done by the system).
$w = -p_{ex}\Delta V$ (for work done against constant external pressure)
If the external pressure is not constant but changes in finite steps, the total work is the sum of work done in each step: $w = -\sum p_{ex} \Delta V$. Graphically, this is the area under the $p_{ex}$ vs $V$ curve.
In a reversible process, the change occurs in infinite steps, and the system is always in near-equilibrium with the surroundings. The external pressure ($p_{ex}$) is always infinitesimally close to the internal pressure of the gas ($p_{in}$). For an expansion, $p_{ex} = p_{in} - dp$; for compression, $p_{ex} = p_{in} + dp$. In the limit of infinitesimal change ($dV$), $dp \rightarrow 0$, so $p_{ex} \approx p_{in}$. Work done in a reversible process is calculated by integrating the pressure-volume work:
$w_{rev} = -\int_{V_i}^{V_f} p_{ex} dV$
Under reversible conditions, $p_{ex} = p_{in}$ (let's call it $p$), so:
$w_{rev} = -\int_{V_i}^{V_f} p \, dV$
For an ideal gas undergoing an isothermal (constant temperature, $T$) reversible process, $p = nRT/V$. Substituting this into the integral:
$w_{rev} = -\int_{V_i}^{V_f} \frac{nRT}{V} \, dV$
Since $n$ and $T$ are constant for an isothermal process:
$w_{rev} = -nRT \int_{V_i}^{V_f} \frac{1}{V} \, dV = -nRT [\ln V]_{V_i}^{V_f} = -nRT (\ln V_f - \ln V_i)$
$w_{rev} = -nRT \ln \frac{V_f}{V_i} = -2.303 nRT \log \frac{V_f}{V_i}$
Free expansion: Expansion against zero external pressure ($p_{ex}=0$, expansion into vacuum). In this case, $w = -p_{ex}\Delta V = 0$. No work is done during free expansion, regardless of whether the process is reversible or irreversible (as there is no opposing pressure). For an ideal gas, internal energy depends only on temperature ($\Delta U = 0$ for isothermal process). If an ideal gas undergoes free expansion (adiabatic, $q=0$), $\Delta U = q+w = 0+0=0$. If it undergoes isothermal free expansion, $\Delta U=0$ (because it's isothermal), and $w=0$ (because it's free expansion), so $q=\Delta U - w = 0 - 0 = 0$. For an ideal gas undergoing isothermal expansion into vacuum, $\Delta U=0$, $w=0$, and $q=0$.
Summarising $\Delta U = q+w$ for different processes involving ideal gas:
- Isothermal irreversible: $\Delta U = 0$, so $q = -w = p_{ex}(V_f - V_i)$.
- Isothermal reversible: $\Delta U = 0$, so $q = -w = nRT \ln(V_f/V_i)$.
- Adiabatic change: $q = 0$, so $\Delta U = w_{ad}$.
Problem 6.2. Two litres of an ideal gas at a pressure of 10 atm expands isothermally at 25 °C into a vacuum until its total volume is 10 litres. How much heat is absorbed and how much work is done in the expansion ?
Answer:
The gas expands into a vacuum. This is a free expansion, where the external pressure ($p_{ex}$) is zero.
Work done ($w$) = $-p_{ex} \Delta V$. Since $p_{ex}=0$, $w = 0 \times (10 \text{ L} - 2 \text{ L}) = 0$ J.
The expansion is isothermal (temperature is constant). For an ideal gas, the internal energy ($\Delta U$) depends only on temperature. Since the temperature is constant, $\Delta U = 0$.
According to the First Law of Thermodynamics: $\Delta U = q + w$.
Substitute the values: $0 = q + 0$, which means $q = 0$ J.
No work is done, and no heat is absorbed or released. Work done is 0 J, and heat absorbed is 0 J.
Problem 6.3. Consider the same expansion, but this time against a constant external pressure of 1 atm.
Answer:
The expansion is against a constant external pressure ($p_{ex} = 1$ atm). The initial volume $V_i = 2$ L and final volume $V_f = 10$ L. The process is isothermal at 25$^\circ$C, and the gas is ideal.
Work done ($w$) = $-p_{ex} \Delta V = -p_{ex} (V_f - V_i)$.
$w = -(1 \text{ atm}) \times (10 \text{ L} - 2 \text{ L}) = -(1 \text{ atm}) \times (8 \text{ L}) = -8 \text{ L atm}$.
To convert L atm to Joules: 1 L atm $\approx$ 101.325 J.
$w = -8 \text{ L atm} \times 101.325 \text{ J/L atm} \approx -810.6$ J.
Since the process is isothermal and the gas is ideal, the internal energy change $\Delta U = 0$.
According to the First Law: $\Delta U = q + w$.
$0 = q + (-8 \text{ L atm})$
$q = 8 \text{ L atm}$.
$q \approx 810.6$ J.
Work done is -8 L atm (-810.6 J) (work done by the system). Heat absorbed is 8 L atm (810.6 J).
Problem 6.4. Consider the expansion given in problem 6.2, for 1 mol of an ideal gas conducted reversibly.
Answer:
The expansion is isothermal (25$^\circ$C, $T = 25+273.15 = 298.15$ K, text uses 298 K, let's use 298 K) and reversible. The gas is ideal, and the amount is $n=1$ mol. Initial volume $V_i = 2$ L, final volume $V_f = 10$ L. Initial pressure $p_1=10$ atm. We can calculate $n$ from the initial conditions using $pV=nRT$ to confirm (though $n=1$ is given). $n = p_1V_i / (RT) = (10 \text{ atm} \times 2 \text{ L}) / (0.08206 \text{ L atm K}^{-1} \text{ mol}^{-1} \times 298 \text{ K}) \approx 0.818$ mol. The problem statement says "for 1 mol", so let's assume $n=1$ mol is the correct amount, overriding the inconsistency with $p_1V_1T_1$. Let's proceed with $n=1$ mol.
For an isothermal reversible expansion of an ideal gas:
Work done ($w_{rev}$) = $-nRT \ln \frac{V_f}{V_i} = -2.303 nRT \log \frac{V_f}{V_i}$.
Using $n=1$ mol, $R = 0.08206$ L atm K$^{-1}$ mol$^{-1}$, $T = 298$ K, $V_f=10$ L, $V_i=2$ L:
$w_{rev} = -2.303 \times (1 \text{ mol}) \times (0.08206 \text{ L atm K}^{-1} \text{ mol}^{-1}) \times (298 \text{ K}) \times \log (\frac{10 \text{ L}}{2 \text{ L}})$.
$w_{rev} = -2.303 \times 0.08206 \times 298 \times \log(5)$ L atm.
$w_{rev} = -2.303 \times 0.08206 \times 298 \times 0.69897$ L atm.
$w_{rev} \approx -39.45$ L atm.
To convert L atm to Joules: 1 L atm $\approx$ 101.325 J.
$w_{rev} \approx -39.45 \text{ L atm} \times 101.325 \text{ J/L atm} \approx -3997$ J $\approx -4.00$ kJ.
Since the process is isothermal and the gas is ideal, the internal energy change $\Delta U = 0$.
According to the First Law: $\Delta U = q + w$.
$0 = q + (-39.45 \text{ L atm})$.
$q = 39.45 \text{ L atm}$.
$q \approx 4000$ J $\approx 4.00$ kJ.
Work done is approximately -39.45 L atm (-4.00 kJ). Heat absorbed is approximately 39.45 L atm (4.00 kJ).
Enthalpy, H
While constant volume conditions ($V=constant, \Delta V=0, w=0, \Delta U=q_V$) are used in bomb calorimeters, most chemical reactions occur at constant pressure (like atmospheric pressure). To quantify heat changes at constant pressure, we define a new state function called Enthalpy ($H$).
Enthalpy ($H$) is defined as the sum of the internal energy ($U$) and the pressure-volume product ($pV$).
$H = U + pV$
Since $U$, $p$, and $V$ are state functions, $H$ is also a state function; its value depends only on the state of the system.
Consider a change in state at constant pressure:
$\Delta H = \Delta U + \Delta(pV)$
At constant pressure ($p$ is constant):
$\Delta H = \Delta U + p\Delta V$
From the First Law, $\Delta U = q + w$. At constant pressure, work done is $w = -p_{ex}\Delta V$. If the external pressure is equal to the constant internal pressure ($p_{ex}=p$), then $w = -p\Delta V$.
Substituting $w = -p\Delta V$ into $\Delta U = q+w$ gives $\Delta U = q + (-p\Delta V)$, or $\Delta U = q - p\Delta V$.
Rearranging gives $q = \Delta U + p\Delta V$.
Comparing this with the definition of enthalpy change at constant pressure, $\Delta H = \Delta U + p\Delta V$, we see that:
$\Delta H = q_p$
The change in enthalpy ($\Delta H$) is equal to the heat absorbed by the system at constant pressure ($q_p$). Therefore, $\Delta H$ is also path-independent under constant pressure conditions. For exothermic reactions, heat is released by the system, $q_p$ is negative, so $\Delta H$ is negative. For endothermic reactions, heat is absorbed by the system, $q_p$ is positive, so $\Delta H$ is positive.
For reactions involving gases, the volume change ($\Delta V$) at constant temperature and pressure can be related to the change in the number of moles of gas ($\Delta n_g$).
Let the total volume of gaseous reactants be $V_A$ ($n_A$ moles) and the total volume of gaseous products be $V_B$ ($n_B$ moles). At constant $p$ and $T$, using the ideal gas law:
$pV_A = n_A RT$
$pV_B = n_B RT$}
$p(V_B - V_A) = (n_B - n_A) RT$
$p\Delta V = \Delta n_g RT$
where $\Delta n_g = (n_B - n_A) = (\text{moles of gaseous products}) - (\text{moles of gaseous reactants})$.
Substituting this into the enthalpy change equation $\Delta H = \Delta U + p\Delta V$ gives:
$\Delta H = \Delta U + \Delta n_g RT$
This equation is useful for converting between $\Delta H$ and $\Delta U$ for reactions involving gases.
Problem 6.5. If water vapour is assumed to be a perfect gas, molar enthalpy change for vapourisation of 1 mol of water at 1bar and 100°C is 41kJ mol–1. Calculate the internal energy change, when 1 mol of water is vapourised at 1 bar pressure and 100°C.
Answer:
The process is the vaporisation of 1 mol of water at constant pressure (1 bar) and temperature (100$^\circ$C). We are given the molar enthalpy of vaporisation ($\Delta_{vap} H$).
Process: $\text{H}_2\text{O(l)} \rightarrow \text{H}_2\text{O(g)}$ (for 1 mol)
Given: $\Delta_{vap} H = 41$ kJ mol$^{-1}$. Pressure $p = 1$ bar. Temperature $T = 100^\circ$C = $100 + 273.15 = 373.15$ K. (Text uses 373 K, let's use 373 K).
We need to calculate the internal energy change ($\Delta_{vap} U$). The relationship between $\Delta H$ and $\Delta U$ is $\Delta H = \Delta U + \Delta n_g RT$.
First, calculate $\Delta n_g$ for the reaction: $\text{H}_2\text{O(l)} \rightarrow \text{H}_2\text{O(g)}$.
Number of moles of gaseous products = 1 (for H$_2$O(g)).
Number of moles of gaseous reactants = 0 (H$_2$O(l) is liquid).
$\Delta n_g = n_{products(gas)} - n_{reactants(gas)} = 1 - 0 = 1$ mol.
Now, rearrange the equation to find $\Delta U$: $\Delta U = \Delta H - \Delta n_g RT$.
Substitute the values. $R = 8.314$ J K$^{-1}$ mol$^{-1}$. Since $\Delta H$ is in kJ, we should use $R$ in kJ K$^{-1}$ mol$^{-1}$ or convert $\Delta H$ to J. Let's convert $R$: $R = 8.314 \times 10^{-3}$ kJ K$^{-1}$ mol$^{-1}$.
$\Delta_{vap} U = 41 \text{ kJ mol}^{-1} - (1 \text{ mol}) \times (8.314 \times 10^{-3} \text{ kJ K}^{-1} \text{ mol}^{-1}) \times (373 \text{ K})$.
$\Delta_{vap} U = 41 \text{ kJ mol}^{-1} - (1 \times 8.314 \times 10^{-3} \times 373) \text{ kJ mol}^{-1}$.
$\Delta_{vap} U = 41 \text{ kJ mol}^{-1} - 3.099 \text{ kJ mol}^{-1}$. (Text gets 3.096 kJ mol-1)
$\Delta_{vap} U \approx 37.901 \text{ kJ mol}^{-1}$. (Text gets 37.904 kJ mol-1)
The internal energy change is approximately 37.90 kJ mol$^{-1}$.
Heat Capacity
When heat is transferred to a system, its temperature usually rises. The amount of heat ($q$) required to change the temperature of a substance by a certain amount ($\Delta T$) is related by the heat capacity ($C$).
$q = C \Delta T$
Heat capacity depends on the amount of substance, its composition, and the process conditions (constant volume or constant pressure).
- Molar heat capacity ($C_m$ or $C$): Heat capacity for one mole of the substance ($C_m = C/n$). It is the heat required to raise the temperature of one mole by one degree Celsius (or Kelvin). Units: J mol$^{-1}$ K$^{-1}$.
- Specific heat capacity ($c$ or $s$): Heat capacity for one unit mass of the substance. It is the heat required to raise the temperature of one unit mass by one degree Celsius (or Kelvin). Units: J g$^{-1}$ K$^{-1}$.
Total heat ($q$) required for a sample of mass ($m$) with specific heat ($c$) is $q = m \times c \times \Delta T$. This is also equal to $C \Delta T$ if $C$ is the heat capacity for the entire sample.
The Relationship Between Cp And Cv For An Ideal Gas
Heat capacity measured at constant volume is denoted as $C_V$. Heat capacity measured at constant pressure is denoted as $C_p$.
At constant volume: $\Delta U = q_V = C_V \Delta T$
At constant pressure: $\Delta H = q_p = C_p \Delta T$
For an ideal gas, the relationship between $\Delta H$ and $\Delta U$ is $\Delta H = \Delta U + \Delta n_g RT$. For a change involving one mole of ideal gas ($\Delta n_g = 1$ for processes like $A(s) \rightarrow B(g)$ where B is 1 mole, or considering $C_p$ and $C_V$ per mole), $\Delta H = \Delta U + R\Delta T$.
Substituting the expressions for $\Delta H$ and $\Delta U$ in terms of heat capacities:
$C_p \Delta T = C_V \Delta T + R \Delta T$
Dividing by $\Delta T$ (for a non-zero temperature change):
$C_p = C_V + R$
or
$C_p - C_V = R$
This relationship holds for one mole of an ideal gas, where $R$ is the universal gas constant. It shows that $C_p$ is greater than $C_V$ because at constant pressure, some energy is used to do expansion work against the surroundings as the temperature rises, whereas at constant volume, all the energy goes into increasing the internal energy (and thus temperature).
Measurement Of Du And Dh: Calorimetry
Calorimetry is the experimental technique used to measure the heat changes associated with chemical or physical processes. It involves carrying out the process in a device called a calorimeter, which is typically immersed in a liquid (like water) with a known heat capacity. By measuring the temperature change of the liquid and calorimeter, the heat evolved or absorbed can be determined.
Energy changes are usually measured under two specific conditions: constant volume or constant pressure.
Du Measurements
Heat change at constant volume ($q_V$) is measured using a bomb calorimeter. This apparatus consists of a sealed, rigid steel vessel (the "bomb") where the reaction takes place, immersed in a water bath within an insulated container. The constant volume of the bomb ensures that no PV work is done ($\Delta V=0$, $w=0$).
For combustion reactions, a known amount of the substance is placed in the bomb with excess oxygen and ignited. The heat released by the reaction ($q_V$) is transferred to the water bath and the calorimeter itself, causing a temperature rise ($\Delta T$). The heat absorbed by the calorimeter ($q_{calorimeter}$) is calculated using the heat capacity of the calorimeter system ($C_{calorimeter}$) and the observed temperature change:
$q_{calorimeter} = C_{calorimeter} \Delta T$
Since the system (reaction) and the surroundings (calorimeter + water) are isolated, the heat released by the reaction is equal in magnitude but opposite in sign to the heat absorbed by the calorimeter: $q_{reaction} = -q_{calorimeter}$.
As the reaction is carried out at constant volume, the heat measured ($q_V$) is equal to the change in internal energy of the reaction:
$\Delta_r U = q_V = -q_{calorimeter} = -C_{calorimeter} \Delta T$
The heat capacity of the calorimeter is usually determined by a separate calibration experiment (e.g., burning a substance with known heat of combustion or using electrical heating).
Dh Measurements
Heat change at constant pressure ($q_p$) is typically measured using a calorimeter open to the atmosphere or maintained at a constant external pressure. This is simpler than a bomb calorimeter and is often done in a coffee-cup calorimeter or similar apparatus.
In this case, the reaction occurs at constant pressure (usually atmospheric pressure). The heat absorbed or evolved during the reaction ($q_p$) causes a temperature change ($\Delta T$) in the solution or system. Assuming the heat capacity of the system ($C_{system}$) is known, the heat absorbed or evolved is calculated:
$q_p = C_{system} \Delta T$
At constant pressure, the heat measured ($q_p$) is equal to the change in enthalpy of the reaction:
$\Delta_r H = q_p = C_{system} \Delta T$
For an exothermic reaction, heat is evolved ($q_p < 0$), so $\Delta_r H < 0$. For an endothermic reaction, heat is absorbed ($q_p > 0$), so $\Delta_r H > 0$.
The heat capacity of the system (e.g., solution) can be estimated from the specific heat capacity and mass of the components.
Problem 6.6. 1g of graphite is burnt in a bomb calorimeter in excess of oxygen at 298 K and 1 atmospheric pressure according to the equation C (graphite) + O$_2$ (g) $\rightarrow$ CO$_2$ (g) During the reaction, temperature rises from 298 K to 299 K. If the heat capacity of the bomb calorimeter is 20.7kJ/K, what is the enthalpy change for the above reaction at 298 K and 1 atm?
Answer:
The reaction is combustion of graphite in a bomb calorimeter, which measures the heat change at constant volume ($q_V$).
Given mass of graphite = 1 g.
Reaction: C (graphite) + O$_2$ (g) $\rightarrow$ CO$_2$ (g).
Initial temperature $T_1 = 298$ K, Final temperature $T_2 = 299$ K. Temperature change $\Delta T = T_2 - T_1 = 299 \text{ K} - 298 \text{ K} = 1$ K.
Heat capacity of the bomb calorimeter ($C_{calorimeter}$) = 20.7 kJ/K.
The heat absorbed by the calorimeter is $q_{calorimeter} = C_{calorimeter} \Delta T$.
$q_{calorimeter} = (20.7 \text{ kJ/K}) \times (1 \text{ K}) = 20.7$ kJ.
The heat released by the reaction ($q_V$) is equal in magnitude but opposite in sign to the heat absorbed by the calorimeter.
$q_V = -q_{calorimeter} = -20.7$ kJ.
Since the process is carried out at constant volume in a bomb calorimeter, the heat measured is the change in internal energy for the reaction involving 1 g of graphite.
$\Delta U$ (for 1 g graphite) = $q_V = -20.7$ kJ.
We are asked for the enthalpy change ($\Delta H$) for the reaction as written in the equation (which involves 1 mole of graphite, C). First, calculate $\Delta U$ per mole.
Molar mass of graphite (C) = 12.011 g/mol. (Let's use 12.0 g/mol as is common in basic problems).
$\Delta_r U$ (for 1 mol graphite) = $\Delta U$ (for 1 g) $\times$ Molar mass of C
$\Delta_r U = (-20.7 \text{ kJ/g}) \times (12.0 \text{ g/mol}) = -248.4$ kJ mol$^{-1}$. (Text gets -248 x 10^2 kJ mol-1, which seems a typo, it should be -2.48 x 10^2 kJ mol-1)
Now, we need to find the enthalpy change ($\Delta_r H$) for the reaction at 298 K and 1 atm. We use the relationship $\Delta H = \Delta U + \Delta n_g RT$.
The reaction is: C (graphite) + O$_2$ (g) $\rightarrow$ CO$_2$ (g).
Calculate $\Delta n_g$ = (moles of gaseous products) - (moles of gaseous reactants).
Gaseous products: 1 mole of CO$_2$ (g).
Gaseous reactants: 1 mole of O$_2$ (g). (Graphite C(s) is solid, so it's not included in $\Delta n_g$).
$\Delta n_g = 1 - 1 = 0$ moles.
So, $\Delta_r H = \Delta_r U + (0) RT = \Delta_r U$.
Therefore, the enthalpy change for the reaction at 298 K and 1 atm is equal to the internal energy change calculated at constant volume.
$\Delta_r H = -248.4$ kJ mol$^{-1}$.
The enthalpy change for the reaction at 298 K and 1 atm is approximately -248.4 kJ mol$^{-1}$.
(Note: The discrepancy with the text's final value (-2.48 x 10^2 kJ mol-1) might be due to significant figure conventions or slight variations in molar mass of C used (12.0 vs 12.011) or R value if used in intermediate steps - though here $\Delta n_g=0$. The most likely issue is the text's calculation resulting in 2.48 instead of 248.4 from -20.7 * 12.0). Let's use the text's intermediate value as a check. If $\Delta_r U = -2.48 \times 10^2$ kJ mol$^{-1}$, then $\Delta_r H = -2.48 \times 10^2$ kJ mol$^{-1} = -248$ kJ mol$^{-1}$. My calculation -248.4 is close. I'll state my calculated value.)
Enthalpy Change, Drh Of A Reaction – Reaction Enthalpy
The enthalpy change accompanying a chemical reaction, where reactants are converted into products, is called the reaction enthalpy ($\Delta_r H$). It is defined as the difference between the sum of the enthalpies of the products and the sum of the enthalpies of the reactants, taking into account their stoichiometric coefficients in the balanced chemical equation.
$\Delta_r H = \sum a_i H_{products} - \sum b_i H_{reactants}$
where $a_i$ and $b_i$ are the stoichiometric coefficients of the products and reactants, respectively, and $H$ represents the molar enthalpy of each substance.
For an exothermic reaction, heat is released, and the products have lower enthalpy than the reactants, so $\Delta_r H$ is negative. For an endothermic reaction, heat is absorbed, and the products have higher enthalpy than the reactants, so $\Delta_r H$ is positive.
Reaction enthalpy is a crucial quantity for understanding the energy balance of chemical processes.
Standard Enthalpy Of Reactions
To compare reaction enthalpies consistently, standard conditions are defined. The standard enthalpy of reaction ($\Delta_r H^0$) is the enthalpy change for a reaction when all reactants and products are in their standard states at a specified temperature (usually 298 K). The standard state of a substance is its most stable form at 1 bar pressure and the specified temperature. For example, the standard state of oxygen at 298 K is O$_2$(g) at 1 bar; for carbon, it's graphite C(graphite) at 1 bar.
Standard conditions are indicated by the superscript $^0$ (e.g., $\Delta H^0$).
Enthalpy Changes During Phase Transformations
Changes in the physical state of a substance (phase transformations) also involve enthalpy changes, typically occurring at constant temperature and pressure.
- Standard Enthalpy of Fusion ($\Delta_{fus} H^0$): The enthalpy change when one mole of a solid melts into its liquid form at its melting point and at standard pressure (1 bar). Melting is endothermic ($\Delta_{fus} H^0 > 0$).
- Standard Enthalpy of Vaporization ($\Delta_{vap} H^0$): The enthalpy change when one mole of a liquid vaporizes into its gaseous form at its boiling point and at standard pressure (1 bar). Vaporization is endothermic ($\Delta_{vap} H^0 > 0$).
- Standard Enthalpy of Sublimation ($\Delta_{sub} H^0$): The enthalpy change when one mole of a solid sublimes directly into its gaseous form at a constant temperature and at standard pressure (1 bar). Sublimation is endothermic ($\Delta_{sub} H^0 > 0$).
For a substance, the enthalpy of sublimation is approximately the sum of the enthalpy of fusion and the enthalpy of vaporization at the same temperature: $\Delta_{sub} H^0 \approx \Delta_{fus} H^0 + \Delta_{vap} H^0$.
The magnitude of these phase transition enthalpies reflects the strength of intermolecular forces. Substances with stronger intermolecular forces (like water with hydrogen bonding) have higher $\Delta_{fus} H^0$ and $\Delta_{vap} H^0$ values compared to substances with weaker forces (like acetone with weaker dipole-dipole interactions).
| Substance | $\Delta_{fus}H^0$ (kJ mol$^{-1}$) | $T_f$ (K) | $\Delta_{vap}H^0$ (kJ mol$^{-1}$) | $T_b$ (K) |
| H$_2$O | 6.00 | 273.15 | 40.79 | 373.15 |
| NaCl | 28.8 | 1081 | 170.0 | 1665 |
| C$_2$H$_5$OH | 4.93 | 159.0 | 38.56 | 351.5 |
| CCl$_4$ | 2.5 | 250.0 | 30.4 | 350.0 |
Problem 6.7. A swimmer coming out from a pool is covered with a film of water weighing about 18g. How much heat must be supplied to evaporate this water at 298 K ? Calculate the internal energy of vaporisation at 298K. Dvap H0 for water at 298K= 44.01kJ mol–1
Answer:
We need to calculate the heat required to evaporate 18 g of water at 298 K and the internal energy change for this process.
Given mass of water = 18 g.
Molar mass of water (H$_2$O) = $2 \times 1.008 + 16.00 = 18.016$ g/mol. (Using 18 g/mol as in text examples).
Number of moles of water = $\frac{\text{Mass}}{\text{Molar mass}} = \frac{18 \text{ g}}{18 \text{ g/mol}} = 1$ mol.
Given standard enthalpy of vaporisation of water at 298 K ($\Delta_{vap} H^0$) = 44.01 kJ mol$^{-1}$. This is the heat required to vaporize 1 mole of water at 298 K and standard pressure (1 bar).
Since we have 1 mol of water, the heat supplied to evaporate 18g (1 mol) of water at 298 K is equal to $\Delta_{vap} H^0$.
Heat supplied ($q_p$) = $1 \text{ mol} \times 44.01 \text{ kJ mol}^{-1} = 44.01$ kJ.
Now, calculate the internal energy change of vaporisation ($\Delta_{vap} U$) at 298 K. The process is $\text{H}_2\text{O(l)} \rightarrow \text{H}_2\text{O(g)}$.
We use the relationship $\Delta H = \Delta U + \Delta n_g RT$.
Calculate $\Delta n_g$: $\Delta n_g = n_{products(gas)} - n_{reactants(gas)} = 1 \text{ mol (H$_2$O(g))} - 0 \text{ mol (H$_2$O(l))} = 1$ mol.
Rearrange to find $\Delta U$: $\Delta U = \Delta H - \Delta n_g RT$.
Substitute the values. $\Delta H = 44.01$ kJ mol$^{-1}$. $\Delta n_g = 1$ mol. $T = 298$ K. $R = 8.314$ J K$^{-1}$ mol$^{-1} = 8.314 \times 10^{-3}$ kJ K$^{-1}$ mol$^{-1}$.
$\Delta_{vap} U = 44.01 \text{ kJ mol}^{-1} - (1 \text{ mol}) \times (8.314 \times 10^{-3} \text{ kJ K}^{-1} \text{ mol}^{-1}) \times (298 \text{ K})$.
$\Delta_{vap} U = 44.01 \text{ kJ mol}^{-1} - (1 \times 8.314 \times 10^{-3} \times 298) \text{ kJ mol}^{-1}$.
$\Delta_{vap} U = 44.01 \text{ kJ mol}^{-1} - 2.478 \text{ kJ mol}^{-1}$. (Text gets 2.48 kJ mol-1)
$\Delta_{vap} U \approx 41.532 \text{ kJ mol}^{-1}$. (Text gets 41.53 kJ mol-1)
The heat supplied is 44.01 kJ, and the internal energy of vaporisation at 298 K is approximately 41.53 kJ mol$^{-1}$.
Problem 6.8. Assuming the water vapour to be a perfect gas, calculate the internal energy change when 1 mol of water at 100°C and 1 bar pressure is converted to ice at 0°C. Given the enthalpy of fusion of ice is 6.00 kJ mol-1 heat capacity of water is 4.2 J/g°C
Answer:
The overall process is converting 1 mol of H$_2$O(g) at 100$^\circ$C and 1 bar to 1 mol of H$_2$O(s) (ice) at 0$^\circ$C. We need to find the total internal energy change ($\Delta U$) for this process. We can break this into steps where enthalpy changes are known or calculable and then use the $\Delta H = \Delta U + \Delta n_g RT$ relationship for each step or the overall process.
Let's use steps involving known phase transitions and heat capacity:
Step 1: Cool 1 mol H$_2$O(g) from 100$^\circ$C to 100$^\circ$C liquid (condensation).
H$_2$O(g, 100$^\circ$C, 1 bar) $\rightarrow$ H$_2$O(l, 100$^\circ$C, 1 bar). The enthalpy change is the negative of the enthalpy of vaporization at 100$^\circ$C (boiling point at 1 atm $\approx$ 1 bar).
$\Delta H_1 = -\Delta_{vap} H^0$ (at 100$^\circ$C). From Table 6.1, $\Delta_{vap} H^0$ at 373.15 K (100$^\circ$C) is 40.79 kJ mol$^{-1}$.
$\Delta H_1 = -40.79$ kJ mol$^{-1}$.
For this step, $\Delta n_g = n_{gas} - n_{liquid} = 0 - 1 = -1$ mol (gas to liquid). We are told water vapour is a perfect gas. The volume of liquid is negligible compared to gas.
$\Delta U_1 = \Delta H_1 - \Delta n_g RT_1 = -40.79 \text{ kJ mol}^{-1} - (-1 \text{ mol}) \times (8.314 \times 10^{-3} \text{ kJ K}^{-1} \text{ mol}^{-1}) \times (373 \text{ K})$.
$\Delta U_1 = -40.79 \text{ kJ mol}^{-1} + 3.099 \text{ kJ mol}^{-1} \approx -37.691$ kJ mol$^{-1}$.
Step 2: Cool 1 mol H$_2$O(l) from 100$^\circ$C to 0$^\circ$C.
H$_2$O(l, 100$^\circ$C) $\rightarrow$ H$_2$O(l, 0$^\circ$C). This occurs at constant pressure (1 bar). The enthalpy change is $q_p$.
$\Delta H_2 = n \times C_p(\text{liquid}) \times \Delta T$. We are given specific heat capacity $c = 4.2$ J/g$^\circ$C. Molar mass of water is 18 g/mol. Molar heat capacity $C_p(\text{liquid}) = c \times M = 4.2 \text{ J g}^{-1} \text{ ^\circ C}^{-1} \times 18 \text{ g mol}^{-1} = 75.6$ J mol$^{-1}$ $^\circ$C$^{-1} = 75.6$ J mol$^{-1}$ K$^{-1}$.
Temperature change $\Delta T = T_{final} - T_{initial} = 0^\circ\text{C} - 100^\circ\text{C} = -100^\circ\text{C} = -100$ K.
$\Delta H_2 = (1 \text{ mol}) \times (75.6 \text{ J mol}^{-1} \text{ K}^{-1}) \times (-100 \text{ K}) = -7560$ J mol$^{-1} = -7.56$ kJ mol$^{-1}$.
For this step, the substance is liquid, so volume change is negligible. $p\Delta V \approx 0$, so $\Delta U_2 \approx \Delta H_2 = -7.56$ kJ mol$^{-1}$.
Step 3: Freeze 1 mol H$_2$O(l) at 0$^\circ$C to 1 mol H$_2$O(s) at 0$^\circ$C.
H$_2$O(l, 0$^\circ$C) $\rightarrow$ H$_2$O(s, 0$^\circ$C). This occurs at constant temperature and pressure. The enthalpy change is the negative of the enthalpy of fusion at 0$^\circ$C.
$\Delta H_3 = -\Delta_{fus} H^0$ (at 0$^\circ$C). Given $\Delta_{fus} H^0 = 6.00$ kJ mol$^{-1}$ at 0$^\circ$C.
$\Delta H_3 = -6.00$ kJ mol$^{-1}$.
For this step (liquid to solid), volume change is negligible. $p\Delta V \approx 0$, so $\Delta U_3 \approx \Delta H_3 = -6.00$ kJ mol$^{-1}$.
Total internal energy change for the overall process:
$\Delta U_{\text{total}} = \Delta U_1 + \Delta U_2 + \Delta U_3$
$\Delta U_{\text{total}} \approx (-37.691 \text{ kJ mol}^{-1}) + (-7.56 \text{ kJ mol}^{-1}) + (-6.00 \text{ kJ mol}^{-1})$
$\Delta U_{\text{total}} \approx -51.251$ kJ mol$^{-1}$.
Let's check the text's approach using total enthalpy change first.
Total enthalpy change $\Delta H_{\text{total}} = \Delta H_1 + \Delta H_2 + \Delta H_3$
$\Delta H_{\text{total}} = (-40.79 \text{ kJ mol}^{-1}) + (-7.56 \text{ kJ mol}^{-1}) + (-6.00 \text{ kJ mol}^{-1}) = -54.35$ kJ mol$^{-1}$.
Overall process: H$_2$O(g, 100$^\circ$C, 1 bar) $\rightarrow$ H$_2$O(s, 0$^\circ$C, 1 bar). For the overall process, the initial state is gas and the final state is solid. The volume change $\Delta V_{\text{total}}$ is significant only due to the initial gas phase. We can approximate $\Delta V_{\text{total}} \approx V_{gas} - V_{solid} \approx V_{gas}$ (since $V_{solid}$ is negligible compared to $V_{gas}$).
For the initial gas state: $p_1V_1 = nRT_1 \implies V_1 = nRT_1/p_1 = (1 \text{ mol}) \times (0.08314 \text{ bar L K}^{-1} \text{ mol}^{-1}) \times (373 \text{ K}) / (1 \text{ bar}) \approx 31.0$ L.
For the final solid state, volume is very small, let's approximate $V_2 \approx 0$.
$\Delta V_{\text{total}} = V_2 - V_1 \approx 0 - 31.0 \text{ L} = -31.0$ L.
$p \Delta V_{\text{total}} \approx (1 \text{ bar}) \times (-31.0 \text{ L}) = -31.0$ L bar. Convert to kJ: 1 L bar = $100$ J $= 0.1$ kJ.
$p \Delta V_{\text{total}} \approx -31.0 \times 0.1 \text{ kJ} = -3.1$ kJ.
Relationship for overall process: $\Delta H_{\text{total}} = \Delta U_{\text{total}} + p\Delta V_{\text{total}}$ (since pressure is constant).
$\Delta U_{\text{total}} = \Delta H_{\text{total}} - p\Delta V_{\text{total}} \approx -54.35 \text{ kJ} - (-3.1 \text{ kJ}) = -54.35 + 3.1 \approx -51.25$ kJ.
This matches my previous calculation of $\Delta U_{\text{total}}$ by summing individual $\Delta U$ steps. The text's steps seem to be calculating total $\Delta H$ and then using $\Delta U = \Delta H$ assuming $p\Delta V \approx 0$ based on the final state (solid), which is incorrect because the initial state is gas with significant volume. The text example calculation seems flawed or based on simplified intermediate steps not fully shown.
Using the text's approximate values for $\Delta H_1, \Delta H_2, \Delta H_3$ from their problem wording description: $\Delta H_1$ (cool vapour 100C to liquid 100C) = $-\Delta_{vap}H$ = -41 kJ/mol (from Pr. 6.5). $\Delta H_2$ (cool liquid 100C to 0C) = $-7.56$ kJ/mol. $\Delta H_3$ (freeze liquid 0C to ice 0C) = $-6.00$ kJ/mol.
Total $\Delta H = -41 - 7.56 - 6.00 = -54.56$ kJ/mol.
Overall process: H$_2$O(g, 100$^\circ$C, 1 bar) $\rightarrow$ H$_2$O(s, 0$^\circ$C, 1 bar).
$\Delta n_g$ for overall process is $0$ (solid) - $1$ (gas) = $-1$ mol.
$T_{initial} = 373$ K, $T_{final} = 273$ K.
The relationship $\Delta H = \Delta U + \Delta n_g RT$ is for a process at *constant* temperature. Applying it to a process with temperature change needs integration or considering the overall process as changing from one state to another.
Let's calculate $\Delta U$ for each step based on the text's provided $\Delta H$ values for the steps and $T$:
- Step 1: Gas (100C) $\rightarrow$ Liquid (100C). $\Delta H_1 = -41$ kJ/mol. $\Delta n_g = 0-1 = -1$ mol. $\Delta U_1 = \Delta H_1 - \Delta n_g RT = -41 - (-1)(8.314 \times 10^{-3})(373) = -41 + 3.099 \approx -37.901$ kJ/mol.
- Step 2: Liquid (100C) $\rightarrow$ Liquid (0C). $\Delta H_2 = -7.56$ kJ/mol. $\Delta n_g \approx 0$. $\Delta U_2 \approx \Delta H_2 = -7.56$ kJ/mol.
- Step 3: Liquid (0C) $\rightarrow$ Solid (0C). $\Delta H_3 = -6.00$ kJ/mol. $\Delta n_g \approx 0$. $\Delta U_3 \approx \Delta H_3 = -6.00$ kJ/mol.
Total $\Delta U = \Delta U_1 + \Delta U_2 + \Delta U_3 = -37.901 - 7.56 - 6.00 \approx -51.461$ kJ/mol.
This is closer to -51.25 I got earlier. The text result -13.56 kJ mol-1 for the whole process as $\Delta U$ appears incorrect given the steps and starting point (vapour at 100C). The text seems to imply $\Delta H = -13.56$ based on Step 2+3, which is the cooling and freezing of liquid from 100C, but the starting point is vapour at 100C. The text example seems to omit the vaporisation step enthalpy.
Let's calculate total $\Delta U$ by considering the overall process. $\Delta U$ is a state function, so we just need initial and final states.
Initial state: 1 mol H$_2$O(g) at 100$^\circ$C (373 K), 1 bar.
Final state: 1 mol H$_2$O(s) at 0$^\circ$C (273 K), 1 bar.
This requires using absolute enthalpies or statistical mechanics, which is beyond the scope of this unit. However, if we assume $\Delta H$ and $\Delta U$ relationship holds for the overall process involving gas:
Overall $\Delta H = \Delta H_1 + \Delta H_2 + \Delta H_3 = -40.79 - 7.56 - 6.00 = -54.35$ kJ/mol (using the correct $\Delta_{vap}H$ at 100C).
Overall $\Delta n_g$ = $0$ (solid) - $1$ (gas) = $-1$ mol.
$\Delta H_{\text{total}} = \Delta U_{\text{total}} + \Delta n_g RT_{\text{avg}}$ ? No, the temperature is not constant.
The correct relationship is $\Delta U = \Delta H - \Delta(pV)$. For constant pressure, $\Delta U = \Delta H - p\Delta V$. $p\Delta V = p(V_{final} - V_{initial})$. $V_{final}$ (solid) is negligible. $V_{initial}$ (gas) $= nRT_{initial}/p = 1 \times R \times 373 / 1$.
$\Delta U = \Delta H - p(0 - nRT_{initial}/p) = \Delta H + nRT_{initial}$.
$\Delta U = -54.35 \text{ kJ mol}^{-1} + (1 \text{ mol}) \times (8.314 \times 10^{-3} \text{ kJ K}^{-1} \text{ mol}^{-1}) \times (373 \text{ K})$.
$\Delta U = -54.35 + 3.099 \approx -51.251$ kJ mol$^{-1}$.
There appears to be a significant error in the text's example calculation. Based on standard thermodynamic principles and given data, the calculated $\Delta U$ is around -51.25 kJ/mol.
However, if the question *meant* to start with liquid water at 100$^\circ$C, then $\Delta H = \Delta H_2 + \Delta H_3 = -7.56 - 6.00 = -13.56$ kJ/mol. Since $\Delta n_g = 0$ for liquid/solid transitions, $\Delta U \approx \Delta H = -13.56$ kJ/mol. This matches the text's final result but contradicts the initial state given as "water vapour at 100$^\circ$C". Assuming the question is stated correctly, the text's answer is wrong. Assuming the text's answer is correct means the question is stated incorrectly (should start with liquid at 100C).
Let's stick to the question as written and my derivation based on it. However, I should point out the inconsistency.
Based on the question as written (vapour at 100$^\circ$C to ice at 0$^\circ$C) and the provided data, the total internal energy change is approximately -51.25 kJ mol$^{-1}$. If the question meant liquid water at 100$^\circ$C to ice at 0$^\circ$C, the internal energy change would be approximately -13.56 kJ mol$^{-1}$. Given the text's final answer matches the latter case, it is highly likely the question has a typo and should say "1 mol of water at 100°C and 1 bar pressure is converted to ice at 0°C." implying liquid state initially.
Assuming the question meant H$_2$O(l, 100$^\circ$C, 1 bar) $\rightarrow$ H$_2$O(s, 0$^\circ$C, 1 bar):
Steps are: Liquid (100C) $\rightarrow$ Liquid (0C) $\rightarrow$ Solid (0C).
$\Delta H = \Delta H_2 + \Delta H_3 = -7.56 \text{ kJ mol}^{-1} + (-6.00 \text{ kJ mol}^{-1}) = -13.56$ kJ mol$^{-1}$.
For processes involving only liquid and solid, $\Delta n_g = 0$, so $\Delta U \approx \Delta H$.
$\Delta U \approx -13.56$ kJ mol$^{-1}$.
Given the strong possibility of a typo in the source text, I will present the answer for the *likely intended* question (liquid to solid) as it matches the text's provided numerical result, but will add a note about the original wording.
Assuming the question meant converting 1 mol of water liquid at 100$^\circ$C and 1 bar pressure to ice at 0$^\circ$C:
Step 1: Cool 1 mol H$_2$O(l) from 100$^\circ$C to 0$^\circ$C.
$\Delta H_1 = 1 \text{ mol} \times C_p(\text{liquid}) \times \Delta T = 1 \text{ mol} \times (75.6 \text{ J mol}^{-1} \text{ K}^{-1}) \times (-100 \text{ K}) = -7560$ J = -7.56 kJ.
Step 2: Freeze 1 mol H$_2$O(l) at 0$^\circ$C to H$_2$O(s) at 0$^\circ$C.
$\Delta H_2 = -\Delta_{fus} H^0 = -6.00$ kJ mol$^{-1}$.
Total enthalpy change $\Delta H = \Delta H_1 + \Delta H_2 = -7.56 \text{ kJ} + (-6.00 \text{ kJ}) = -13.56$ kJ.
For processes involving only liquid and solid phases, the volume change is generally small, so $p\Delta V \approx 0$. Thus, $\Delta U \approx \Delta H$.
The internal energy change is approximately -13.56 kJ mol$^{-1}$.
Standard Enthalpy Of Formation
The standard molar enthalpy of formation ($\Delta_f H^0$) of a compound is the standard enthalpy change for the reaction in which one mole of the compound is formed from its constituent elements in their most stable states of aggregation (reference states) at a specified temperature (usually 298 K) and 1 bar pressure.
The reference state of an element is its pure form in its most stable physical state at 1 bar and the specified temperature (e.g., O$_2$(g) for oxygen, C(graphite, s) for carbon, Br$_2$(l) for bromine, H$_2$(g) for hydrogen).
By convention, the standard enthalpy of formation of an element in its reference state is zero ($\Delta_f H^0 = 0$).
Example formation reactions and their standard enthalpies of formation:
- Formation of liquid water: $\text{H}_2\text{(g)} + \frac{1}{2}\text{O}_2\text{(g)} \rightarrow \text{H}_2\text{O(l)}$; $\Delta_f H^0 = -285.8$ kJ mol$^{-1}$
- Formation of methane gas: $\text{C(graphite, s)} + 2\text{H}_2\text{(g)} \rightarrow \text{CH}_4\text{(g)}$; $\Delta_f H^0 = -74.81$ kJ mol$^{-1}$
Note that $\Delta_f H^0$ refers to the formation of exactly one mole of the product compound from elements in their standard states.
Standard enthalpies of formation are useful for calculating the standard enthalpy of any reaction ($\Delta_r H^0$). The general equation for this calculation is based on Hess's Law:
$\Delta_r H^0 = \sum n_p \Delta_f H^0 (\text{products}) - \sum n_r \Delta_f H^0 (\text{reactants})$
where $n_p$ and $n_r$ are the stoichiometric coefficients of the products and reactants in the balanced equation.
| Substance | $\Delta_fH^0$ (kJ mol$^{-1}$) | Substance | $\Delta_fH^0$ (kJ mol$^{-1}$) |
| H$_2$O(l) | –285.83 | CO(g) | –110.53 |
| H$_2$O(g) | –241.82 | CO$_2$(g) | –393.51 |
| HF(g) | –273.3 | CH$_4$(g) | –74.81 |
| HCl(g) | –92.32 | C$_2$H$_6$(g) | –84.68 |
| HBr(g) | –36.40 | C$_3$H$_8$(g) | –103.85 |
| HI(g) | 26.48 | CH$_3$OH(l) | –238.86 |
| NaCl(s) | –384.1 | C$_2$H$_5$OH(l) | –277.69 |
| KCl(s) | –436.7 | C$_6$H$_6$(l) | 49.04 |
| KBr(s) | –393.8 | Glucose(s) | –1260 |
| KI(s) | –327.0 | Al$_2$O$_3$(s) | –1675.7 |
| AgCl(s) | –127.07 | Fe$_2$O$_3$(s) | –824.2 |
| AgBr(s) | –100.84 | CaCO$_3$(s) | –1206.9 |
| AgI(s) | –55.92 | CaO(s) | –635.1 |
| SO$_2$(g) | –296.8 | ||
| SO$_3$(g) | –395.7 |
Example: Decomposition of CaCO$_3$(s) at standard conditions:
CaCO$_3$(s) $\rightarrow$ CaO(s) + CO$_2$(g); $\Delta_r H^0 = ?$
Using the values from Table 6.2:
$\Delta_r H^0 = [\Delta_f H^0(\text{CaO, s}) + \Delta_f H^0(\text{CO}_2\text{, g})] - [\Delta_f H^0(\text{CaCO}_3\text{, s})]$
$\Delta_r H^0 = [(-635.1 \text{ kJ mol}^{-1}) + (-393.5 \text{ kJ mol}^{-1})] - [-1206.9 \text{ kJ mol}^{-1}]$
$\Delta_r H^0 = (-1028.6 + 1206.9) \text{ kJ mol}^{-1} = 178.3 \text{ kJ mol}^{-1}$.
This positive value confirms that the decomposition of CaCO$_3$(s) is an endothermic reaction.
Thermochemical Equations
A thermochemical equation is a balanced chemical equation that includes the physical states of all reactants and products (including allotropic forms for elements) and the standard enthalpy change ($\Delta_r H^0$) for the reaction as written.
Example: $\text{C}_2\text{H}_5\text{OH(l)} + 3\text{O}_2\text{(g)} \rightarrow 2\text{CO}_2\text{(g)} + 3\text{H}_2\text{O(l)}$; $\Delta_r H^0 = -1367$ kJ mol$^{-1}$.
Conventions for thermochemical equations:
- Coefficients refer to the number of moles of reactants and products.
- The $\Delta_r H^0$ value corresponds to the reaction as written (per mole of reaction). The unit kJ mol$^{-1}$ refers to 'per mole of reaction', where one mole of reaction corresponds to the stoichiometric amounts shown. Doubling the coefficients doubles the $\Delta_r H^0$ value.
- Reversing the equation reverses the sign of $\Delta_r H^0$.
Hess’s Law Of Constant Heat Summation
Based on enthalpy being a state function, Hess's Law states that the total enthalpy change for a reaction is the same, whether the reaction is carried out in one step or in a series of steps. If a reaction can be expressed as the sum of several steps, the standard enthalpy change for the overall reaction is the sum of the standard enthalpy changes for the individual steps at the same temperature.
If Reaction (Overall): A $\rightarrow$ B with $\Delta_r H$
Can occur via steps: A $\rightarrow$ C ($\Delta_r H_1$), C $\rightarrow$ D ($\Delta_r H_2$), D $\rightarrow$ B ($\Delta_r H_3$)
Then, $\Delta_r H = \Delta_r H_1 + \Delta_r H_2 + \Delta_r H_3$.
Hess's law is extremely useful because it allows us to calculate enthalpy changes for reactions that are difficult or impossible to measure directly by manipulating thermochemical equations (reversing, multiplying by coefficients, adding them). For example, to find $\Delta_r H^0$ for C(graphite, s) + $\frac{1}{2}$O$_2$(g) $\rightarrow$ CO(g), which is hard to measure directly, we can use the known enthalpies of combustion of C and CO to CO$_2$ and combine them using Hess's Law.
(i) $\text{C(graphite, s)} + \text{O}_2\text{(g)} \rightarrow \text{CO}_2\text{(g)}$; $\Delta_r H^0_1 = -393.5$ kJ mol$^{-1}$
(ii) $\text{CO(g)} + \frac{1}{2}\text{O}_2\text{(g)} \rightarrow \text{CO}_2\text{(g)}$; $\Delta_r H^0_2 = -283.0$ kJ mol$^{-1}$
Reverse equation (ii) to get CO as a product: $\text{CO}_2\text{(g)} \rightarrow \text{CO(g)} + \frac{1}{2}\text{O}_2\text{(g)}$; $\Delta_r H^0_3 = +283.0$ kJ mol$^{-1}$
Add equation (i) and (iii):
$\text{C(graphite, s)} + \text{O}_2\text{(g)} + \text{CO}_2\text{(g)} \rightarrow \text{CO}_2\text{(g)} + \text{CO(g)} + \frac{1}{2}\text{O}_2\text{(g)}$
Cancel CO$_2$(g) and $\frac{1}{2}$O$_2$(g) from both sides:
$\text{C(graphite, s)} + \frac{1}{2}\text{O}_2\text{(g)} \rightarrow \text{CO(g)}$
The enthalpy change for this reaction is $\Delta_r H^0 = \Delta_r H^0_1 + \Delta_r H^0_3 = (-393.5 \text{ kJ mol}^{-1}) + (283.0 \text{ kJ mol}^{-1}) = -110.5$ kJ mol$^{-1}$.
Enthalpies For Different Types Of Reactions
Specific names are given to enthalpy changes associated with particular types of chemical reactions or processes.
Standard Enthalpy Of Combustion
The standard enthalpy of combustion ($\Delta_c H^0$) is the standard enthalpy change when one mole of a substance undergoes complete combustion (reacts with excess oxygen) at a specified temperature (usually 298 K) and 1 bar pressure. Combustion reactions are generally exothermic ($\Delta_c H^0 < 0$).
Example: Complete combustion of one mole of methane:
$\text{CH}_4\text{(g)} + 2\text{O}_2\text{(g)} \rightarrow \text{CO}_2\text{(g)} + 2\text{H}_2\text{O(l)}$; $\Delta_c H^0 = -890.3$ kJ mol$^{-1}$.
Standard enthalpies of combustion are useful for calculating enthalpies of formation using Hess's Law.
Problem 6.9. The combustion of one mole of benzene takes place at 298 K and 1 atm. After combustion, CO$_2$(g) and H$_2$O (l) are produced and 3267.0 kJ of heat is liberated. Calculate the standard enthalpy of formation, Df H0 of benzene. Standard enthalpies of formation of CO$_2$(g) and 2 H$_2$O(l) are –393.5 kJ mol–1 and – 285.83 kJ mol–1 respectively.
Answer:
Given: Combustion of 1 mol of benzene (C$_6$H$_6$). Standard conditions (298 K, 1 atm $\approx$ 1 bar). Heat liberated is 3267.0 kJ. Heat liberated means $\Delta_c H^0$ is negative.
Reaction: $\text{C}_6\text{H}_6\text{(l)} + \frac{15}{2}\text{O}_2\text{(g)} \rightarrow 6\text{CO}_2\text{(g)} + 3\text{H}_2\text{O(l)}$; $\Delta_c H^0 = -3267.0$ kJ mol$^{-1}$.
Given standard enthalpies of formation:
- $\Delta_f H^0(\text{CO}_2\text{, g}) = -393.5$ kJ mol$^{-1}$. This corresponds to: $\text{C(graphite)} + \text{O}_2\text{(g)} \rightarrow \text{CO}_2\text{(g)}$
- $\Delta_f H^0(\text{H}_2\text{O, l}) = -285.83$ kJ mol$^{-1}$. This corresponds to: $\text{H}_2\text{(g)} + \frac{1}{2}\text{O}_2\text{(g)} \rightarrow \text{H}_2\text{O(l)}$
We need to calculate the standard enthalpy of formation of liquid benzene ($\Delta_f H^0(\text{C}_6\text{H}_6\text{, l})$). This corresponds to the reaction:
$\text{6C(graphite, s)} + \text{3H}_2\text{(g)} \rightarrow \text{C}_6\text{H}_6\text{(l)}$; $\Delta_f H^0(\text{C}_6\text{H}_6\text{, l}) = ?$
We can use Hess's Law, combining the combustion reaction of benzene and the formation reactions of CO$_2$ and H$_2$O.
The target reaction is: (1) $6\text{C(graphite)} + 3\text{H}_2\text{(g)} \rightarrow \text{C}_6\text{H}_6\text{(l)}$
We have the following reactions with known $\Delta H^0$:
(2) $\text{C}_6\text{H}_6\text{(l)} + \frac{15}{2}\text{O}_2\text{(g)} \rightarrow 6\text{CO}_2\text{(g)} + 3\text{H}_2\text{O(l)}$; $\Delta H^0_2 = -3267.0$ kJ mol$^{-1}$.
(3) $\text{C(graphite)} + \text{O}_2\text{(g)} \rightarrow \text{CO}_2\text{(g)}$; $\Delta H^0_3 = -393.5$ kJ mol$^{-1}$.
(4) $\text{H}_2\text{(g)} + \frac{1}{2}\text{O}_2\text{(g)} \rightarrow \text{H}_2\text{O(l)}$; $\Delta H^0_4 = -285.83$ kJ mol$^{-1}$.
To get the target reaction (1), we need C(graphite) and H$_2$(g) on the reactant side and C$_6$H$_6$(l) on the product side. Also, O$_2$, CO$_2$, and H$_2$O must cancel out.
Take reaction (3) and multiply by 6 (to get 6C):
(5) $6 \times [(3)]$: $6\text{C(graphite)} + 6\text{O}_2\text{(g)} \rightarrow 6\text{CO}_2\text{(g)}$; $\Delta H^0_5 = 6 \times (-393.5) = -2361.0$ kJ.
Take reaction (4) and multiply by 3 (to get 3H$_2$):
(6) $3 \times [(4)]$: $3\text{H}_2\text{(g)} + \frac{3}{2}\text{O}_2\text{(g)} \rightarrow 3\text{H}_2\text{O(l)}$; $\Delta H^0_6 = 3 \times (-285.83) = -857.49$ kJ.
Add reactions (5) and (6):
(7) $6\text{C(graphite)} + 3\text{H}_2\text{(g)} + 6\text{O}_2\text{(g)} + \frac{3}{2}\text{O}_2\text{(g)} \rightarrow 6\text{CO}_2\text{(g)} + 3\text{H}_2\text{O(l)}$
(7) $6\text{C(graphite)} + 3\text{H}_2\text{(g)} + \frac{15}{2}\text{O}_2\text{(g)} \rightarrow 6\text{CO}_2\text{(g)} + 3\text{H}_2\text{O(l)}$; $\Delta H^0_7 = \Delta H^0_5 + \Delta H^0_6 = -2361.0 + (-857.49) = -3218.49$ kJ.
Now, compare reaction (7) with the target reaction (1) and reaction (2). Reaction (7) has the reactants of (1) and the products of (2). Reaction (2) has the reactant of (1) on the product side and products of (7) on the product side.
Let's reverse reaction (2) to get C$_6$H$_6$(l) on the product side:
(8) Reverse [(2)]: $6\text{CO}_2\text{(g)} + 3\text{H}_2\text{O(l)} \rightarrow \text{C}_6\text{H}_6\text{(l)} + \frac{15}{2}\text{O}_2\text{(g)}$; $\Delta H^0_8 = - \Delta H^0_2 = +3267.0$ kJ.
Add reaction (7) and reaction (8):
$6\text{C} + 3\text{H}_2 + \frac{15}{2}\text{O}_2 + 6\text{CO}_2 + 3\text{H}_2\text{O} \rightarrow 6\text{CO}_2 + 3\text{H}_2\text{O} + \text{C}_6\text{H}_6 + \frac{15}{2}\text{O}_2$
Cancel terms common to both sides ($6\text{CO}_2\text{(g)}$, $3\text{H}_2\text{O(l)}$, $\frac{15}{2}\text{O}_2\text{(g)}$):
$6\text{C(graphite)} + 3\text{H}_2\text{(g)} \rightarrow \text{C}_6\text{H}_6\text{(l)}$
This is the target reaction (1). The enthalpy change is the sum of $\Delta H^0$ for reactions (7) and (8).
$\Delta H^0_1 = \Delta H^0_7 + \Delta H^0_8 = (-3218.49 \text{ kJ}) + (3267.0 \text{ kJ}) = 48.51$ kJ.
The standard enthalpy of formation of benzene is +48.51 kJ mol$^{-1}$.
Enthalpy Of Atomization
The standard enthalpy of atomization ($\Delta_a H^0$) is the standard enthalpy change when one mole of a substance is completely broken down into individual gaseous atoms. This process involves breaking all the bonds in the substance and converting the elements into their atomic gaseous state.
Example: Atomization of methane gas:
$\text{CH}_4\text{(g)} \rightarrow \text{C(g)} + 4\text{H(g)}$; $\Delta_a H^0 = 1665$ kJ mol$^{-1}$.
Example: Atomization of sodium metal (solid to gaseous atoms):
$\text{Na(s)} \rightarrow \text{Na(g)}$; $\Delta_a H^0 = 108.4$ kJ mol$^{-1}$. For metals, this is also the enthalpy of sublimation.
Bond Enthalpy
Chemical reactions involve breaking existing bonds in reactants and forming new bonds in products. Energy is required to break bonds (endothermic), and energy is released when new bonds form (exothermic).
For diatomic molecules, the bond dissociation enthalpy is the enthalpy change when one mole of a specific type of bond in a gaseous diatomic molecule is broken to form gaseous atoms. This is the same as the enthalpy of atomization for a diatomic molecule.
Example: $\text{H}_2\text{(g)} \rightarrow 2\text{H(g)}$; $\Delta_{H-H} H^0 = 435.0$ kJ mol$^{-1}$.
For polyatomic molecules, the energy required to break a specific type of bond can vary slightly depending on the molecule and the location of the bond within the molecule. For example, the four C-H bonds in methane (CH$_4$) require different amounts of energy to break successively.
In such cases, the concept of mean bond enthalpy is used. It is the average value of bond dissociation enthalpies for a given type of bond across different molecules or in successive bond breaking steps within the same molecule.
For CH$_4$, the mean C-H bond enthalpy is calculated by dividing the total atomization enthalpy by the number of C-H bonds (4):
Mean $\Delta_{C-H} H^0 = \frac{\Delta_a H^0(\text{CH}_4)}{4} = \frac{1665 \text{ kJ mol}^{-1}}{4} \approx 416$ kJ mol$^{-1}$.
Mean bond enthalpy values can be used to estimate the enthalpy change for a gas phase reaction using the relationship:
$\Delta_r H^0 \approx \sum (\text{bond enthalpies of reactants}) - \sum (\text{bond enthalpies of products})$
This equation represents the energy required to break all bonds in the reactants minus the energy released when all bonds in the products are formed. Note this is an approximation, especially if the reaction is not entirely in the gas phase or if mean bond enthalpies are used.
| H | C | N | O | F | Si | P | S | Cl | Br | I | |
|---|---|---|---|---|---|---|---|---|---|---|---|
| H | 435.8 | 414 | 389 | 464 | 569 | 293 | 318 | 339 | 431 | 368 | 297 |
| C | 347 | 293 | 351 | 439 | 289 | 264 | 259 | 330 | 276 | 238 | |
| N | 159 | 201 | 272 | - | 209 | - | 201 | 243 | - | ||
| O | 138 | 184 | 368 | 351 | - | 205 | - | 201 | |||
| F | 155 | 540 | 490 | 327 | 255 | 197 | - | ||||
| Si | 176 | 213 | 226 | 360 | 289 | 213 | |||||
| P | 213 | 230 | 331 | 272 | 213 | ||||||
| S | 213 | 251 | 213 | - | |||||||
| Cl | 243 | 218 | 209 | ||||||||
| Br | 192 | 180 | |||||||||
| I | 151 |
| Bond | Enthalpy (kJ mol$^{-1}$) |
| N = N | 418 |
| C = C | 611 |
| O = O | 498 |
| N $\equiv$ N | 946 |
| C $\equiv$ C | 837 |
| C = N | 615 |
| C = O | 741 |
| C $\equiv$ N | 891 |
| C $\equiv$ O | 1070 |
Lattice Enthalpy
The Lattice Enthalpy ($\Delta_{lattice}H^0$) of an ionic compound is the standard enthalpy change when one mole of the solid ionic compound completely dissociates into its constituent gaseous ions.
Example: $\text{Na}^+\text{Cl}^-\text{(s)} \rightarrow \text{Na}^+\text{(g)} + \text{Cl}^-\text{(g)}$; $\Delta_{lattice}H^0 = +788$ kJ mol$^{-1}$.
Lattice enthalpy is a measure of the strength of the electrostatic forces holding the ions together in the crystal lattice. It is difficult to measure directly. It is usually determined indirectly using a Born-Haber cycle, which is an enthalpy diagram based on Hess's Law.
A Born-Haber cycle links the enthalpy of formation of an ionic compound to various steps, including atomization of the elements, ionization enthalpy of the metal, electron gain enthalpy of the non-metal, and lattice enthalpy.
Summing the enthalpy changes around the cycle allows calculation of the unknown lattice enthalpy (e.g., $\Delta_f H^0(\text{NaCl, s}) = \Delta_{sub} H^0(\text{Na, s}) + \Delta_i H^0(\text{Na, g}) + \frac{1}{2}\Delta_{bond} H^0(\text{Cl}_2\text{, g}) + \Delta_{eg} H^0(\text{Cl, g}) + \Delta_{lattice} H^0(\text{NaCl, s})$). By convention, the cycle usually adds steps from the elements in their standard states to the gaseous ions, and then from the gaseous ions to the solid ionic compound (lattice enthalpy, which is exothermic from ions to solid, or endothermic from solid to ions).
$\Delta_{lattice}H^0$ from solid to gaseous ions is typically a large positive value.
Enthalpy Of Solution
The enthalpy of solution ($\Delta_{sol}H^0$) is the standard enthalpy change when one mole of a solute dissolves in a specified amount of solvent. The enthalpy of solution at infinite dilution refers to dissolving in so much solvent that interactions between solute particles are negligible.
When an ionic compound dissolves in water, it involves two main steps:
- Breaking the crystal lattice: This requires energy, equal to the lattice enthalpy ($\Delta_{lattice}H^0$, positive).
- Hydrating the ions: Gaseous ions are surrounded by water molecules, releasing energy (enthalpy of hydration, $\Delta_{hyd}H^0$, negative).
The enthalpy of solution is the sum of these enthalpy changes:
$\Delta_{sol}H^0 = \Delta_{lattice}H^0 + \Delta_{hyd}H^0$ (for ionic compounds dissolving in water)
If $\Delta_{hyd}H^0$ is more negative than $\Delta_{lattice}H^0$ is positive, the dissolution is exothermic ($\Delta_{sol}H^0 < 0$). If $\Delta_{lattice}H^0$ is larger, the dissolution is endothermic ($\Delta_{sol}H^0 > 0$). For many salts, $\Delta_{sol}H^0$ is positive, and solubility increases with temperature.
Enthalpy Of Dilution
The enthalpy of dilution is the enthalpy change that occurs when a solution of a particular concentration is diluted by adding more solvent. This process involves separating solute particles and solvent molecules and forming new interactions upon dilution. It depends on the initial concentration and the amount of solvent added.
Example: Dissolving HCl gas in water.
$\text{HCl(g)} + 10 \text{ aq.} \rightarrow \text{HCl.10 aq.}$ $\Delta H = -69.01$ kJ/mol
$\text{HCl(g)} + 40 \text{ aq.} \rightarrow \text{HCl.40 aq.}$ $\Delta H = -72.79$ kJ/mol
The enthalpy change for diluting HCl solution from containing 10 mol water to 40 mol water is the difference between these two steps, effectively representing $\text{HCl.10 aq.} + 30 \text{ aq.} \rightarrow \text{HCl.40 aq.}$. Using Hess's Law, we reverse the first reaction and add the second:
$\text{HCl.10 aq.} \rightarrow \text{HCl(g)} + 10 \text{ aq.}$ $\Delta H = +69.01$ kJ/mol
$\text{HCl(g)} + 40 \text{ aq.} \rightarrow \text{HCl.40 aq.}$ $\Delta H = -72.79$ kJ/mol
Adding them: $\text{HCl.10 aq.} + 30 \text{ aq.} \rightarrow \text{HCl.40 aq.}$ $\Delta H = 69.01 - 72.79 = -3.78$ kJ/mol.
The enthalpy of dilution from 10 aq to 40 aq is -3.78 kJ/mol.
Spontaneity
The First Law of Thermodynamics deals with energy conservation but does not predict the direction of a process. Many naturally occurring processes (like heat flow from hot to cold, gases expanding) proceed spontaneously in one direction and do not reverse on their own. Spontaneity in thermodynamics refers to the potential of a process to occur without external assistance. It doesn't imply the speed of the process. A spontaneous process is generally irreversible; reversing it requires external intervention.
Is Decrease In Enthalpy A Criterion For Spontaneity ?
Many spontaneous processes, like a ball rolling downhill or the combustion of fuel, are exothermic (release energy, $\Delta H < 0$). This might suggest that a decrease in enthalpy is the driving force for spontaneity.
Indeed, most exothermic reactions are spontaneous (e.g., formation of NH$_3$, HCl, H$_2$O from elements). However, there are spontaneous processes that are endothermic ($\Delta H > 0$), such as the dissolution of many salts (e.g., KNO$_3$ in water makes the water colder) or the reaction:
$\frac{1}{2}\text{N}_2\text{(g)} + \text{O}_2\text{(g)} \rightarrow \text{NO}_2\text{(g)}$; $\Delta_r H^0 = +33.2$ kJ mol$^{-1}$ (spontaneous endothermic)
Since spontaneous endothermic reactions exist, a decrease in enthalpy ($\Delta H < 0$) is a driving factor for spontaneity but not the sole criterion.
Entropy And Spontaneity
What drives spontaneous processes that don't involve a decrease in enthalpy, or even involve an increase? Consider the spontaneous diffusion of two gases when a partition is removed.
Initially, the gases are ordered (A on one side, B on the other). After diffusion, they are mixed; the system is less ordered, more random, or more chaotic. This increase in disorder is a characteristic of spontaneous processes in isolated systems.
This leads to the introduction of another thermodynamic state function: Entropy ($S$). Entropy is a measure of the degree of randomness or disorder in a system. Higher disorder corresponds to higher entropy. For a given substance, $S_{solid} < S_{liquid} < S_{gas}$.
Entropy ($S$), like $U$ and $H$, is a state function ($\Delta S$ is path-independent). The change in entropy for a reversible process ($\Delta S_{sys}$) is defined as the heat transferred reversibly ($q_{rev}$) divided by the absolute temperature ($T$):
$\Delta S_{sys} = \frac{q_{sys, rev}}{T}$
The Second Law of Thermodynamics provides a criterion for spontaneity based on entropy: For any spontaneous process, the total entropy of the universe (system + surroundings) increases.
$\Delta S_{total} = \Delta S_{sys} + \Delta S_{surr} > 0$ (for a spontaneous process)
At equilibrium, $\Delta S_{total} = 0$. A process is non-spontaneous if $\Delta S_{total} < 0$.
Thus, an increase in the total entropy of the universe is the driving force for spontaneous processes.
Problem 6.10. Predict in which of the following, entropy increases/decreases :
(i) A liquid crystallizes into a solid.
(ii) Temperature of a crystalline solid is raised from 0 K to 115 K.
(iii) NaHCO$_3$(s) $\rightarrow$ Na$_2$CO$_3$(s) + CO$_2$(g) + H$_2$O(l)
(iv) H$_2$(g) $\rightarrow$ 2 H(g)
Answer:
Entropy is a measure of disorder or randomness. We need to assess the change in the degree of disorder in each process.
(i) A liquid crystallizes into a solid: In the solid state, particles are arranged in a highly ordered, fixed structure compared to the liquid state where particles are still close but can move past each other. Ordering increases during crystallisation. Therefore, entropy decreases.
(ii) Temperature of a crystalline solid is raised from 0 K to 115 K: At 0 K, a perfect crystalline solid has minimum possible entropy (perfect order). As temperature increases, the particles gain kinetic energy and vibrate more vigorously about their lattice positions, increasing disorder. Therefore, entropy increases.
(iii) NaHCO$_3$(s) $\rightarrow$ Na$_2$CO$_3$(s) + CO$_2$(g) + H$_2$O(l): The reactant is a solid, which is an ordered state. The products include one solid, one gas (CO$_2$), and one liquid (H$_2$O). Gases have significantly higher entropy than solids or liquids because their particles are spread out and move randomly. The formation of a gas from a solid represents a large increase in disorder. Therefore, entropy increases.
(iv) H$_2$(g) $\rightarrow$ 2 H(g): One mole of diatomic hydrogen molecules (a gas) is converted into two moles of individual hydrogen atoms (also a gas). The number of independent particles increases from 1 to 2. More particles mean more ways to distribute energy and occupy space, leading to greater disorder. Therefore, entropy increases.
Problem 6.11. For oxidation of iron, 4 Fe(s) + 3 O$_2$(g) $\rightarrow$ 2 Fe$_2$O$_3$(s), entropy change ($\Delta_r S^0$) is – 549.4 JK–1mol–1at 298 K. Inspite of negative entropy change of this reaction, why is the reaction spontaneous? (DrH0for this reaction is –1648 × 10$^3$ J mol–1)
Answer:
A reaction is spontaneous if the total entropy change of the universe ($\Delta S_{total}$) is positive, not just the entropy change of the system ($\Delta S_{sys}$). The total entropy change is the sum of the entropy change of the system and the entropy change of the surroundings: $\Delta S_{total} = \Delta S_{sys} + \Delta S_{surr}$.
Given for the system (the reaction):
- $\Delta S_{sys} = -549.4$ J K$^{-1}$ mol$^{-1}$ at $T = 298$ K. The negative sign indicates the system becomes more ordered (gas reactants forming a solid product).
- $\Delta_r H^0 = -1648 \times 10^3$ J mol$^{-1}$ = -1648 kJ mol$^{-1}$. The reaction is highly exothermic, releasing heat to the surroundings.
Now, calculate the entropy change of the surroundings ($\Delta S_{surr}$). When heat is transferred from the system to the surroundings, the surroundings absorb heat equal to the negative of the enthalpy change of the system (at constant pressure, $q_{surr} = -\Delta H_{sys}$). The entropy change of the surroundings is given by $\Delta S_{surr} = \frac{q_{surr}}{T} = \frac{-\Delta H_{sys}}{T}$.
$\Delta S_{surr} = \frac{-(-1648 \times 10^3 \text{ J mol}^{-1})}{298 \text{ K}} = \frac{1648000 \text{ J mol}^{-1}}{298 \text{ K}} \approx 5529.53$ J K$^{-1}$ mol$^{-1}$.
Finally, calculate the total entropy change ($\Delta S_{total}$).
$\Delta S_{total} = \Delta S_{sys} + \Delta S_{surr} = (-549.4 \text{ J K}^{-1} \text{ mol}^{-1}) + (5529.53 \text{ J K}^{-1} \text{ mol}^{-1})$.
$\Delta S_{total} \approx 4980.13$ J K$^{-1}$ mol$^{-1}$.
Since $\Delta S_{total}$ is positive ($>0$), the process is spontaneous, even though the entropy of the system decreases. The large release of heat to the surroundings increases the disorder of the surroundings significantly, outweighing the decrease in disorder of the system.
The reaction is spontaneous because the entropy increase in the surroundings due to the release of heat is greater than the entropy decrease in the system.
Gibbs Energy And Spontaneity
The total entropy change ($\Delta S_{total}$) is the criterion for spontaneity for the universe. However, calculating $\Delta S_{surr}$ (which involves heat exchange) is not always convenient, especially for closed or open systems. A new thermodynamic function, the Gibbs energy ($G$), is introduced to predict spontaneity based on the properties of the system alone (at constant temperature and pressure).
Gibbs energy ($G$) is defined as:
$G = H - TS$
$G$ is an extensive property and a state function. For a process occurring at constant temperature ($T$), the change in Gibbs energy ($\Delta G$) is related to the changes in enthalpy and entropy of the system by the Gibbs equation:
$\Delta G = \Delta H - T\Delta S$ (at constant $T$)
If the process also occurs at constant pressure, then $\Delta G_{sys} = \Delta H_{sys} - T\Delta S_{sys}$. Dropping the 'sys' subscript, $\Delta G = \Delta H - T\Delta S$.
The relationship between $\Delta G$ and spontaneity at constant pressure and temperature is:
- If $\Delta G < 0$: The process is spontaneous.
- If $\Delta G > 0$: The process is non-spontaneous.
- If $\Delta G = 0$: The system is at equilibrium.
$\Delta G$ represents the maximum amount of non-PV work (useful work) that can be obtained from a process at constant temperature and pressure. Hence, it's also known as free energy.
The Gibbs equation $\Delta G = \Delta H - T\Delta S$ shows how enthalpy, entropy, and temperature influence spontaneity:
- If $\Delta H < 0$ (exothermic) and $\Delta S > 0$ (increase in disorder): $\Delta G = (-) - T(+)$ = negative at all temperatures. Process is always spontaneous.
- If $\Delta H > 0$ (endothermic) and $\Delta S < 0$ (decrease in disorder): $\Delta G = (+) - T(-)$ = positive at all temperatures. Process is never spontaneous.
- If $\Delta H < 0$ (exothermic) and $\Delta S < 0$ (decrease in disorder): $\Delta G = (-) - T(-)$. $\Delta G$ is negative if $| \Delta H | > |T\Delta S |$. Spontaneous at low temperatures.
- If $\Delta H > 0$ (endothermic) and $\Delta S > 0$ (increase in disorder): $\Delta G = (+) - T(+)$. $\Delta G$ is negative if $| T\Delta S | > |\Delta H |$. Spontaneous at high temperatures.
This is why many endothermic reactions that increase disorder become spontaneous at high temperatures.
Entropy And Second Law Of Thermodynamics
As discussed, the Second Law of Thermodynamics states that for any spontaneous process, the total entropy of the universe increases ($\Delta S_{total} > 0$). This explains why spontaneous exothermic reactions (which increase surroundings' entropy) are common, and why processes naturally move towards greater disorder.
Absolute Entropy And Third Law Of Thermodynamics
Particles in a substance have various types of motion (translational, rotational, vibrational). Entropy increases with temperature as these motions become more vigorous. The Third Law of Thermodynamics establishes a reference point for entropy:
The entropy of any pure crystalline substance approaches zero as the temperature approaches absolute zero (0 K).
At absolute zero, a perfect crystal has maximum order, and its entropy is considered zero. This law allows for the calculation of absolute entropy values for pure substances by integrating heat capacity data from 0 K up to a desired temperature. Standard molar entropies ($S^0$) are tabulated, allowing calculation of standard entropy changes for reactions ($\Delta_r S^0 = \sum n_p S^0(\text{products}) - \sum n_r S^0(\text{reactants})$) using a similar approach to Hess's Law for enthalpy.
Gibbs Energy Change And Equilibrium
For a reaction at equilibrium, the forward and reverse processes occur at equal rates, and there is no net change in the system's state. The criterion for equilibrium at constant temperature and pressure using Gibbs energy is:
$\Delta G = 0$ (at equilibrium)
The standard Gibbs energy change for a reaction ($\Delta_r G^0$) is related to the equilibrium constant ($K$) of the reaction. For a reaction $\text{aA} + \text{bB} \rightleftharpoons \text{cC} + \text{dD}$, the equilibrium constant $K$ is given by $K_c = \frac{[C]^c[D]^d}{[A]^a[B]^b}$ (in terms of concentrations) or $K_p = \frac{(p_C)^c(p_D)^d}{(p_A)^a(p_B)^b}$ (in terms of partial pressures for gases). In thermodynamics, $K$ refers to $K_p$ for gas-phase reactions or $K_c$ for solution-phase reactions, often using activities.
The relationship between $\Delta_r G^0$ and $K$ is:
$\Delta_r G^0 = -RT \ln K$
or $\Delta_r G^0 = -2.303 RT \log K$
where $R$ is the gas constant and $T$ is the absolute temperature.
This equation is fundamental as it connects thermodynamics (predicting spontaneity, $\Delta G^0$) with chemical kinetics/equilibrium (predicting the extent of reaction, $K$).
- If $\Delta_r G^0 < 0$, $\ln K$ is positive, $K > 1$. The equilibrium lies towards the products, and the reaction proceeds significantly in the forward direction under standard conditions.
- If $\Delta_r G^0 > 0$, $\ln K$ is negative, $K < 1$. The equilibrium lies towards the reactants, and the reaction does not proceed significantly in the forward direction under standard conditions.
- If $\Delta_r G^0 = 0$, $\ln K = 0$, $K = 1$. The reaction is at equilibrium under standard conditions.
The relationship $\Delta_r G^0 = \Delta_r H^0 - T\Delta_r S^0$ allows us to calculate $\Delta_r G^0$ from standard enthalpy and entropy changes. Then, using $\Delta_r G^0 = -RT \ln K$, we can determine the equilibrium constant $K$ and predict the extent of the reaction.
Problem 6.12. Calculate DrG0 for conversion of oxygen to ozone, 3/2 O$_2$(g) $\rightarrow$ O$_3$(g) at 298 K. if Kp for this conversion is 2.47 × 10–29.
Answer:
We are given the equilibrium constant $K_p$ and the temperature $T$, and asked to calculate the standard Gibbs energy change ($\Delta_r G^0$).
Reaction: $\frac{3}{2}\text{O}_2\text{(g)} \rightarrow \text{O}_3\text{(g)}$
Given: $K_p = 2.47 \times 10^{-29}$. Temperature $T = 298$ K. Gas constant $R = 8.314$ J K$^{-1}$ mol$^{-1}$.
Using the relationship $\Delta_r G^0 = -RT \ln K_p = -2.303 RT \log K_p$.
$\Delta_r G^0 = -2.303 \times (8.314 \text{ J K}^{-1} \text{ mol}^{-1}) \times (298 \text{ K}) \times \log (2.47 \times 10^{-29})$.
Calculate the logarithm: $\log (2.47 \times 10^{-29}) = \log(2.47) + \log(10^{-29}) \approx 0.3927 - 29 = -28.6073$.
$\Delta_r G^0 = -2.303 \times 8.314 \times 298 \times (-28.6073)$ J mol$^{-1}$.
$\Delta_r G^0 \approx -5705.8 \times (-28.6073)$ J mol$^{-1}$.
$\Delta_r G^0 \approx 163290$ J mol$^{-1}$.
Converting to kJ: $\Delta_r G^0 \approx 163.29$ kJ mol$^{-1}$.
Rounding to three significant figures: $\Delta_r G^0 \approx 163$ kJ mol$^{-1}$.
The standard Gibbs energy change for the conversion of oxygen to ozone at 298 K is approximately +163 kJ mol$^{-1}$. The large positive value of $\Delta_r G^0$ and the very small value of $K_p$ indicate that the formation of ozone from oxygen is very non-spontaneous under standard conditions at 298 K.
Problem 6.13. Find out the value of equilibrium constant for the following reaction at 298 K. Standard Gibbs energy change, DrG0 at the given temperature is –13.6 kJ mol–1.
Answer:
We are given the standard Gibbs energy change ($\Delta_r G^0$) and the temperature $T$, and asked to find the equilibrium constant $K$.
Given: $\Delta_r G^0 = -13.6$ kJ mol$^{-1}$. Temperature $T = 298$ K. Gas constant $R = 8.314$ J K$^{-1}$ mol$^{-1}$.
Using the relationship $\Delta_r G^0 = -RT \ln K = -2.303 RT \log K$.
Rearrange the equation to solve for $\log K$:
$\log K = \frac{-\Delta_r G^0}{2.303 RT}$.
Substitute the given values. Convert $\Delta_r G^0$ to Joules: $\Delta_r G^0 = -13.6 \text{ kJ mol}^{-1} = -13.6 \times 10^3 \text{ J mol}^{-1}$.
$\log K = \frac{-(-13.6 \times 10^3 \text{ J mol}^{-1})}{2.303 \times (8.314 \text{ J K}^{-1} \text{ mol}^{-1}) \times (298 \text{ K})}$.
$\log K = \frac{13600}{2.303 \times 8.314 \times 298} = \frac{13600}{5705.8}$.
$\log K \approx 2.3835$. (Text gets 2.38)
Now, find $K$ by taking the antilogarithm:
$K = 10^{\log K} = 10^{2.3835}$.
$K \approx 241.8$. (Text gets 2.4 x 10^2)
Rounding to two significant figures (based on -13.6 kJ/mol, assuming 3 significant figures for 298 K and R):
$K \approx 2.4 \times 10^2$.
The value of the equilibrium constant for the reaction at 298 K is approximately $2.4 \times 10^2$.
Exercises
Question 6.1 Choose the correct answer. A thermodynamic state function is a quantity
(i) used to determine heat changes
(ii) whose value is independent of path
(iii) used to determine pressure volume work
(iv) whose value depends on temperature only.
Answer:
Question 6.2 For the process to occur under adiabatic conditions, the correct condition is:
(i) $\Delta T = 0$
(ii) $\Delta p = 0$
(iii) q = 0
(iv) w = 0
Answer:
Question 6.3 The enthalpies of all elements in their standard states are:
(i) unity
(ii) zero
(iii) < 0
(iv) different for each element
Answer:
Question 6.4. $\Delta U^\ominus$ of combustion of methane is – X kJ mol–1. The value of $\Delta H^\ominus$ is
(i) $= \Delta U^\ominus$
(ii) $> \Delta U^\ominus$
(iii) $< \Delta U^\ominus$
(iv) = 0
Answer:
Question 6.5 The enthalpy of combustion of methane, graphite and dihydrogen at 298 K are, –890.3 kJ $mol^{–1}$ –393.5 kJ $mol^{–1}$, and –285.8 kJ $mol^{–1}$ respectively. Enthalpy of formation of $CH_4(g)$ will be
(i) –74.8 kJ $mol^{–1}$
(ii) –52.27 kJ $mol^{–1}$
(iii) +74.8 kJ $mol^{–1}$
(iv) +52.26 kJ $mol^{–1}$.
Answer:
Question 6.6 A reaction, $A + B \rightarrow C + D + q$ is found to have a positive entropy change. The reaction will be
(i) possible at high temperature
(ii) possible only at low temperature
(iii) not possible at any temperature
(iv) possible at any temperature
Answer:
Question 6.7 In a process, 701 J of heat is absorbed by a system and 394 J of work is done by the system. What is the change in internal energy for the process?
Answer:
Question 6.8 The reaction of cyanamide, $NH_2CN$ (s), with dioxygen was carried out in a bomb calorimeter, and $\Delta U$ was found to be –742.7 kJ $mol^{–1}$ at 298 K. Calculate enthalpy change for the reaction at 298 K.
$NH_2CN(g) + \frac{3}{2} O_2(g) \rightarrow N_2(g) + CO_2(g) + H_2O(l)$
Answer:
Question 6.9 Calculate the number of kJ of heat necessary to raise the temperature of 60.0 g of aluminium from 35°C to 55°C. Molar heat capacity of Al is 24 J $mol^{–1} K^{–1}$.
Answer:
Question 6.10 Calculate the enthalpy change on freezing of 1.0 mol of water at 10.0°C to ice at –10.0°C. $\Delta_{fus}H$ = 6.03 kJ $mol^{–1}$ at 0°C.
$C_p [H_2O(l)]$ = 75.3 J $mol^{–1} K^{–1}$
$C_p [H_2O(s)]$ = 36.8 J $mol^{–1} K^{–1}$
Answer:
Question 6.11 Enthalpy of combustion of carbon to $CO_2$ is –393.5 kJ $mol^{–1}$. Calculate the heat released upon formation of 35.2 g of $CO_2$ from carbon and dioxygen gas.
Answer:
Question 6.12 Enthalpies of formation of $CO(g), CO_2(g), N_2O(g)$ and $N_2O_4(g)$ are –110, – 393, 81 and 9.7 kJ $mol^{–1}$ respectively. Find the value of $\Delta_rH$ for the reaction:
$N_2O_4(g) + 3CO(g) \rightarrow N_2O(g) + 3CO_2(g)$
Answer:
Question 6.13. Given
$N_2(g) + 3H_2(g) \rightarrow 2NH_3(g)$ ; $\Delta_r H^\ominus = –92.4 \text{ kJ mol}^{–1}$
What is the standard enthalpy of formation of $NH_3$ gas?
Answer:
Question 6.14. Calculate the standard enthalpy of formation of $CH_3OH(l)$ from the following data:
$CH_3OH (l) + \frac{3}{2} O_2(g) \rightarrow CO_2(g) + 2H_2O(l)$ ; $\Delta_r H^\ominus = –726 \text{ kJ mol}^{–1}$
$C(\text{graphite}) + O_2(g) \rightarrow CO_2(g)$ ; $\Delta_c H^\ominus = –393 \text{ kJ mol}^{–1}$
$H_2(g) + \frac{1}{2} O_2(g) \rightarrow H_2O(l)$ ; $\Delta_f H^\ominus = –286 \text{ kJ mol}^{–1}$.
Answer:
Question 6.15. Calculate the enthalpy change for the process
$CCl_4(g) \rightarrow C(g) + 4 Cl(g)$
and calculate bond enthalpy of C – Cl in $CCl_4(g)$.
$\Delta_{vap}H^\ominus(CCl_4) = 30.5 \text{ kJ mol}^{–1}$.
$\Delta_f H^\ominus (CCl_4) = –135.5 \text{ kJ mol}^{–1}$.
$\Delta_a H^\ominus (C) = 715.0 \text{ kJ mol}^{–1}$, where $\Delta_a H^\ominus$ is enthalpy of atomisation
$\Delta_a H^\ominus (Cl_2) = 242 \text{ kJ mol}^{–1}$
Answer:
Question 6.16 For an isolated system, $\Delta U = 0$, what will be $\Delta S$ ?
Answer:
Question 6.17 For the reaction at 298 K,
$2A + B \rightarrow C$
$\Delta H = 400 \ kJ \ mol^{–1}$ and $\Delta S = 0.2 \ kJ \ K^{–1} \ mol^{–1}$
At what temperature will the reaction become spontaneous considering $\Delta H$ and $\Delta S$ to be constant over the temperature range.
Answer:
Question 6.18 For the reaction,
$2 Cl(g) \rightarrow Cl_2(g)$, what are the signs of $\Delta H$ and $\Delta S$ ?
Answer:
Question 6.19. For the reaction
$2 A(g) + B(g) \rightarrow 2D(g)$
$\Delta U^\ominus = –10.5 \text{ kJ}$ and $\Delta S^\ominus = –44.1 \text{ JK}^{–1}$.
Calculate $\Delta G^\ominus$ for the reaction, and predict whether the reaction may occur spontaneously.
Answer:
Question 6.20. The equilibrium constant for a reaction is 10. What will be the value of $\Delta G^\ominus$ ? R = 8.314 JK–1 mol–1, T = 300 K.
Answer:
Question 6.21. Comment on the thermodynamic stability of NO(g), given
$\frac{1}{2} N_2(g) + \frac{1}{2} O_2(g) \rightarrow NO(g)$ ; $\Delta_r H^\ominus = 90 \text{ kJ mol}^{–1}$
$NO(g) + \frac{1}{2} O_2(g) \rightarrow NO_2(g)$ : $\Delta_r H^\ominus= –74 \text{ kJ mol}^{–1}$
Answer:
Question 6.22. Calculate the entropy change in surroundings when 1.00 mol of H2O(l) is formed under standard conditions. $\Delta_f H^\ominus = –286 \text{ kJ mol}^{–1}$.
Answer: