Loading Solutions...
Atoms
Physics XII - Chapter 12: Complete NCERT Exercise Solutions
Master Atomic Physics with NCERT solutions for Bohr's model, hydrogen spectrum, energy levels, and atomic transitions.
Question 12.1
Choose the correct alternative from the clues given at the end of each statement:
(a) The size of the atom in Thomson’s model is __________ the atomic size in Rutherford’s model. (much greater than / no different from / much less than.)
(b) In the ground state of __________ electrons are in stable equilibrium, while in __________ electrons always experience a net force. (Thomson’s model / Rutherford’s model.)
(c) A classical atom based on __________ is doomed to collapse. (Thomson’s model / Rutherford’s model.)
(d) An atom has a nearly continuous mass distribution in a __________ but has a highly non-uniform mass distribution in __________. (Thomson’s model / Rutherford’s model.)
(e) The positively charged part of the atom possesses most of the mass in __________. (Rutherford’s model / both the models.)
(a) The size of the atom in Thomson’s model is __________ the atomic size in Rutherford’s model. (much greater than / no different from / much less than.)
(b) In the ground state of __________ electrons are in stable equilibrium, while in __________ electrons always experience a net force. (Thomson’s model / Rutherford’s model.)
(c) A classical atom based on __________ is doomed to collapse. (Thomson’s model / Rutherford’s model.)
(d) An atom has a nearly continuous mass distribution in a __________ but has a highly non-uniform mass distribution in __________. (Thomson’s model / Rutherford’s model.)
(e) The positively charged part of the atom possesses most of the mass in __________. (Rutherford’s model / both the models.)
Answer & Explanation:
(a) no different from
• In both models, atomic size is of the order of 10⁻¹⁰ m.
(b) Thomson’s model / Rutherford’s model
• In Thomson’s model, electrons are embedded in positive charge and are in stable equilibrium.
• In Rutherford’s model, electrons revolve around nucleus and experience centripetal force.
(c) Rutherford’s model
• Accelerating electrons should radiate energy and spiral into nucleus according to classical EM theory.
(d) Thomson’s model / Rutherford’s model
• Thomson: Uniform positive charge distribution.
• Rutherford: Mass concentrated in tiny nucleus.
(e) both the models
• In both models, positive part contains most of the mass.
Complete Answer:
(a) no different from
(b) Thomson’s model, Rutherford’s model
(c) Rutherford’s model
(d) Thomson’s model, Rutherford’s model
(e) both the models
• In both models, atomic size is of the order of 10⁻¹⁰ m.
(b) Thomson’s model / Rutherford’s model
• In Thomson’s model, electrons are embedded in positive charge and are in stable equilibrium.
• In Rutherford’s model, electrons revolve around nucleus and experience centripetal force.
(c) Rutherford’s model
• Accelerating electrons should radiate energy and spiral into nucleus according to classical EM theory.
(d) Thomson’s model / Rutherford’s model
• Thomson: Uniform positive charge distribution.
• Rutherford: Mass concentrated in tiny nucleus.
(e) both the models
• In both models, positive part contains most of the mass.
Complete Answer:
(a) no different from
(b) Thomson’s model, Rutherford’s model
(c) Rutherford’s model
(d) Thomson’s model, Rutherford’s model
(e) both the models
Question 12.2
Suppose you are given a chance to repeat the alpha-particle scattering experiment using a thin sheet of solid hydrogen in place of the gold foil. (Hydrogen is a solid at temperatures below 14 K.) What results do you expect?
Answer & Explanation:
Step 1: Understand the experiment
Rutherford used gold foil (Z = 79, A ≈ 197).
Hydrogen has Z = 1, A = 1.
Step 2: Key factors affecting scattering
• Scattering depends on nuclear charge (Z).
• Scattering angle θ ∝ Z (for given impact parameter).
• Hydrogen nucleus is a single proton (much lighter than gold nucleus).
Step 3: Expected results with hydrogen foil
1. Very small scattering angles: Since Z_H = 1 vs Z_Au = 79, scattering force is 1/79² ≈ 1/6241 times weaker.
2. No large-angle scattering: Probability of backscattering (θ > 90°) would be extremely low.
3. Most alpha particles pass undeviated: Even more than with gold foil.
4. Possible recoil effects: Hydrogen nucleus is light, so alpha particles may transfer significant energy to it.
Conclusion:
The experiment would show negligible large-angle scattering, making it difficult to deduce the nuclear model. Rutherford needed heavy nuclei (like gold) to get observable large-angle scattering.
Rutherford used gold foil (Z = 79, A ≈ 197).
Hydrogen has Z = 1, A = 1.
Step 2: Key factors affecting scattering
• Scattering depends on nuclear charge (Z).
• Scattering angle θ ∝ Z (for given impact parameter).
• Hydrogen nucleus is a single proton (much lighter than gold nucleus).
Step 3: Expected results with hydrogen foil
1. Very small scattering angles: Since Z_H = 1 vs Z_Au = 79, scattering force is 1/79² ≈ 1/6241 times weaker.
2. No large-angle scattering: Probability of backscattering (θ > 90°) would be extremely low.
3. Most alpha particles pass undeviated: Even more than with gold foil.
4. Possible recoil effects: Hydrogen nucleus is light, so alpha particles may transfer significant energy to it.
Conclusion:
The experiment would show negligible large-angle scattering, making it difficult to deduce the nuclear model. Rutherford needed heavy nuclei (like gold) to get observable large-angle scattering.
Question 12.3
What is the shortest wavelength present in the Paschen series of spectral lines?
Answer & Explanation:
Step 1: Paschen series formula
For hydrogen-like atoms: \( \frac{1}{\lambda} = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right) \)
Paschen series: \( n_1 = 3 \), \( n_2 = 4, 5, 6, \ldots \)
Rydberg constant \( R = 1.097 \times 10^7 \text{ m}^{-1} \)
Step 2: Shortest wavelength (series limit)
Shortest λ corresponds to largest energy transition: \( n_2 \to \infty \)
\( \frac{1}{\lambda_{\min}} = R \left( \frac{1}{3^2} - \frac{1}{\infty} \right) = \frac{R}{9} \)
Step 3: Calculation
\( \lambda_{\min} = \frac{9}{R} = \frac{9}{1.097 \times 10^7} \)
\( = 8.204 \times 10^{-7} \text{ m} = 820.4 \text{ nm} \)
Answer: Shortest wavelength in Paschen series = 820 nm (infrared region).
For hydrogen-like atoms: \( \frac{1}{\lambda} = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right) \)
Paschen series: \( n_1 = 3 \), \( n_2 = 4, 5, 6, \ldots \)
Rydberg constant \( R = 1.097 \times 10^7 \text{ m}^{-1} \)
Step 2: Shortest wavelength (series limit)
Shortest λ corresponds to largest energy transition: \( n_2 \to \infty \)
\( \frac{1}{\lambda_{\min}} = R \left( \frac{1}{3^2} - \frac{1}{\infty} \right) = \frac{R}{9} \)
Step 3: Calculation
\( \lambda_{\min} = \frac{9}{R} = \frac{9}{1.097 \times 10^7} \)
\( = 8.204 \times 10^{-7} \text{ m} = 820.4 \text{ nm} \)
Answer: Shortest wavelength in Paschen series = 820 nm (infrared region).
Question 12.4
A difference of 2.3 eV separates two energy levels in an atom. What is the frequency of radiation emitted when the atom makes a transition from the upper level to the lower level?
Answer & Explanation:
Step 1: Energy difference
\( \Delta E = 2.3 \text{ eV} \)
Convert to joules: \( \Delta E = 2.3 \times 1.6 \times 10^{-19} = 3.68 \times 10^{-19} \text{ J} \)
Step 2: Using Planck's relation
\( \Delta E = h\nu \)
\( \nu = \frac{\Delta E}{h} = \frac{3.68 \times 10^{-19}}{6.626 \times 10^{-34}} \)
\( = 5.555 \times 10^{14} \text{ Hz} \)
Answer: Frequency = \( 5.56 \times 10^{14} \text{ Hz} \) (visible light region).
\( \Delta E = 2.3 \text{ eV} \)
Convert to joules: \( \Delta E = 2.3 \times 1.6 \times 10^{-19} = 3.68 \times 10^{-19} \text{ J} \)
Step 2: Using Planck's relation
\( \Delta E = h\nu \)
\( \nu = \frac{\Delta E}{h} = \frac{3.68 \times 10^{-19}}{6.626 \times 10^{-34}} \)
\( = 5.555 \times 10^{14} \text{ Hz} \)
Answer: Frequency = \( 5.56 \times 10^{14} \text{ Hz} \) (visible light region).
Question 12.5
The ground state energy of hydrogen atom is -13.6 eV. What are the kinetic and potential energies of the electron in this state?
Answer & Explanation:
Step 1: For hydrogen atom (Bohr model)
Total energy: \( E = -\frac{13.6}{n^2} \text{ eV} \)
For ground state, \( n = 1 \): \( E = -13.6 \text{ eV} \)
Step 2: Virial theorem for Coulomb potential
For \( F \propto 1/r^2 \): \( K = -E \) and \( U = 2E \)
But careful: In Bohr model specifically:
\( K = -E \) and \( U = 2E \)
Step 3: Calculate
\( K = -E = -(-13.6) = +13.6 \text{ eV} \)
\( U = 2E = 2 \times (-13.6) = -27.2 \text{ eV} \)
Check: \( E = K + U = 13.6 + (-27.2) = -13.6 \text{ eV} \) ✓
Answer:
Kinetic energy = +13.6 eV
Potential energy = -27.2 eV
Total energy: \( E = -\frac{13.6}{n^2} \text{ eV} \)
For ground state, \( n = 1 \): \( E = -13.6 \text{ eV} \)
Step 2: Virial theorem for Coulomb potential
For \( F \propto 1/r^2 \): \( K = -E \) and \( U = 2E \)
But careful: In Bohr model specifically:
\( K = -E \) and \( U = 2E \)
Step 3: Calculate
\( K = -E = -(-13.6) = +13.6 \text{ eV} \)
\( U = 2E = 2 \times (-13.6) = -27.2 \text{ eV} \)
Check: \( E = K + U = 13.6 + (-27.2) = -13.6 \text{ eV} \) ✓
Answer:
Kinetic energy = +13.6 eV
Potential energy = -27.2 eV
Question 12.6
A hydrogen atom initially in the ground level absorbs a photon, which excites it to the \( n = 4 \) level. Determine the wavelength and frequency of the photon.
Answer & Explanation:
Step 1: Energy levels in hydrogen
\( E_n = -\frac{13.6}{n^2} \text{ eV} \)
Ground state (\( n = 1 \)): \( E_1 = -13.6 \text{ eV} \)
\( n = 4 \): \( E_4 = -\frac{13.6}{16} = -0.85 \text{ eV} \)
Step 2: Energy of absorbed photon
\( \Delta E = E_4 - E_1 = (-0.85) - (-13.6) = 12.75 \text{ eV} \)
Step 3: Wavelength from energy
Using \( \Delta E = \frac{hc}{\lambda} \)
\( \lambda = \frac{hc}{\Delta E} \)
Using \( hc = 1240 \text{ eV·nm} \) (convenient):
\( \lambda = \frac{1240}{12.75} = 97.25 \text{ nm} \)
Step 4: Frequency
\( \nu = \frac{c}{\lambda} = \frac{3 \times 10^8}{97.25 \times 10^{-9}} \)
\( = 3.085 \times 10^{15} \text{ Hz} \)
Answer:
Wavelength = 97.3 nm (ultraviolet)
Frequency = \( 3.09 \times 10^{15} \text{ Hz} \)
\( E_n = -\frac{13.6}{n^2} \text{ eV} \)
Ground state (\( n = 1 \)): \( E_1 = -13.6 \text{ eV} \)
\( n = 4 \): \( E_4 = -\frac{13.6}{16} = -0.85 \text{ eV} \)
Step 2: Energy of absorbed photon
\( \Delta E = E_4 - E_1 = (-0.85) - (-13.6) = 12.75 \text{ eV} \)
Step 3: Wavelength from energy
Using \( \Delta E = \frac{hc}{\lambda} \)
\( \lambda = \frac{hc}{\Delta E} \)
Using \( hc = 1240 \text{ eV·nm} \) (convenient):
\( \lambda = \frac{1240}{12.75} = 97.25 \text{ nm} \)
Step 4: Frequency
\( \nu = \frac{c}{\lambda} = \frac{3 \times 10^8}{97.25 \times 10^{-9}} \)
\( = 3.085 \times 10^{15} \text{ Hz} \)
Answer:
Wavelength = 97.3 nm (ultraviolet)
Frequency = \( 3.09 \times 10^{15} \text{ Hz} \)
Question 12.7
(a) Using the Bohr’s model calculate the speed of the electron in a hydrogen atom in the \( n = 1 \), 2, and 3 levels.
(b) Calculate the orbital period in each of these levels.
(b) Calculate the orbital period in each of these levels.
Answer & Explanation:
Step 1: Formulas in Bohr model
Radius: \( r_n = n^2 a_0 \) where \( a_0 = 0.529 \text{ Å} = 5.29 \times 10^{-11} \text{ m} \)
Speed: \( v_n = \frac{v_1}{n} \) where \( v_1 = \frac{e^2}{4\pi\epsilon_0 \hbar} = \frac{c}{137} \) (fine structure constant)
\( v_1 = 2.188 \times 10^6 \text{ m/s} \)
Step 2: (a) Speeds
\( v_1 = 2.188 \times 10^6 \text{ m/s} \)
\( v_2 = \frac{v_1}{2} = 1.094 \times 10^6 \text{ m/s} \)
\( v_3 = \frac{v_1}{3} = 7.293 \times 10^5 \text{ m/s} \)
Step 3: (b) Orbital periods
Period \( T_n = \frac{2\pi r_n}{v_n} \)
\( r_n = n^2 a_0 \)
\( T_n = \frac{2\pi (n^2 a_0)}{(v_1/n)} = \frac{2\pi a_0 n^3}{v_1} \)
Step 4: Calculate periods
\( T_1 = \frac{2\pi \times 5.29 \times 10^{-11}}{2.188 \times 10^6} \)
\( = \frac{3.324 \times 10^{-10}}{2.188 \times 10^6} = 1.519 \times 10^{-16} \text{ s} \)
Since \( T_n \propto n^3 \):
\( T_2 = 2^3 \times T_1 = 8 \times 1.519 \times 10^{-16} = 1.215 \times 10^{-15} \text{ s} \)
\( T_3 = 3^3 \times T_1 = 27 \times 1.519 \times 10^{-16} = 4.101 \times 10^{-15} \text{ s} \)
Answers:
(a) Speeds: \( v_1 = 2.19 \times 10^6 \text{ m/s} \), \( v_2 = 1.09 \times 10^6 \text{ m/s} \), \( v_3 = 7.29 \times 10^5 \text{ m/s} \)
(b) Periods: \( T_1 = 1.52 \times 10^{-16} \text{ s} \), \( T_2 = 1.22 \times 10^{-15} \text{ s} \), \( T_3 = 4.10 \times 10^{-15} \text{ s} \)
Radius: \( r_n = n^2 a_0 \) where \( a_0 = 0.529 \text{ Å} = 5.29 \times 10^{-11} \text{ m} \)
Speed: \( v_n = \frac{v_1}{n} \) where \( v_1 = \frac{e^2}{4\pi\epsilon_0 \hbar} = \frac{c}{137} \) (fine structure constant)
\( v_1 = 2.188 \times 10^6 \text{ m/s} \)
Step 2: (a) Speeds
\( v_1 = 2.188 \times 10^6 \text{ m/s} \)
\( v_2 = \frac{v_1}{2} = 1.094 \times 10^6 \text{ m/s} \)
\( v_3 = \frac{v_1}{3} = 7.293 \times 10^5 \text{ m/s} \)
Step 3: (b) Orbital periods
Period \( T_n = \frac{2\pi r_n}{v_n} \)
\( r_n = n^2 a_0 \)
\( T_n = \frac{2\pi (n^2 a_0)}{(v_1/n)} = \frac{2\pi a_0 n^3}{v_1} \)
Step 4: Calculate periods
\( T_1 = \frac{2\pi \times 5.29 \times 10^{-11}}{2.188 \times 10^6} \)
\( = \frac{3.324 \times 10^{-10}}{2.188 \times 10^6} = 1.519 \times 10^{-16} \text{ s} \)
Since \( T_n \propto n^3 \):
\( T_2 = 2^3 \times T_1 = 8 \times 1.519 \times 10^{-16} = 1.215 \times 10^{-15} \text{ s} \)
\( T_3 = 3^3 \times T_1 = 27 \times 1.519 \times 10^{-16} = 4.101 \times 10^{-15} \text{ s} \)
Answers:
(a) Speeds: \( v_1 = 2.19 \times 10^6 \text{ m/s} \), \( v_2 = 1.09 \times 10^6 \text{ m/s} \), \( v_3 = 7.29 \times 10^5 \text{ m/s} \)
(b) Periods: \( T_1 = 1.52 \times 10^{-16} \text{ s} \), \( T_2 = 1.22 \times 10^{-15} \text{ s} \), \( T_3 = 4.10 \times 10^{-15} \text{ s} \)
Question 12.8
The radius of the innermost electron orbit of a hydrogen atom is \( 5.3 \times 10^{-11} \) m. What are the radii of the \( n = 2 \) and \( n = 3 \) orbits?
Answer & Explanation:
Step 1: Bohr radius formula
\( r_n = n^2 r_1 \) where \( r_1 = 5.3 \times 10^{-11} \text{ m} \)
Step 2: Calculate
For \( n = 2 \): \( r_2 = 2^2 \times r_1 = 4 \times 5.3 \times 10^{-11} \)
\( = 2.12 \times 10^{-10} \text{ m} \)
For \( n = 3 \): \( r_3 = 3^2 \times r_1 = 9 \times 5.3 \times 10^{-11} \)
\( = 4.77 \times 10^{-10} \text{ m} \)
Answer:
\( r_2 = \) \( 2.12 \times 10^{-10} \text{ m} \)**
\( r_3 = \) \( 4.77 \times 10^{-10} \text{ m} \)
\( r_n = n^2 r_1 \) where \( r_1 = 5.3 \times 10^{-11} \text{ m} \)
Step 2: Calculate
For \( n = 2 \): \( r_2 = 2^2 \times r_1 = 4 \times 5.3 \times 10^{-11} \)
\( = 2.12 \times 10^{-10} \text{ m} \)
For \( n = 3 \): \( r_3 = 3^2 \times r_1 = 9 \times 5.3 \times 10^{-11} \)
\( = 4.77 \times 10^{-10} \text{ m} \)
Answer:
\( r_2 = \) \( 2.12 \times 10^{-10} \text{ m} \)**
\( r_3 = \) \( 4.77 \times 10^{-10} \text{ m} \)
Question 12.9
A 12.5 eV electron beam is used to bombard gaseous hydrogen at room temperature. What series of wavelengths will be emitted?
Answer & Explanation:
Step 1: Hydrogen energy levels
\( E_n = -\frac{13.6}{n^2} \text{ eV} \)
\( E_1 = -13.6 \), \( E_2 = -3.4 \), \( E_3 = -1.51 \), \( E_4 = -0.85 \), \( E_5 = -0.544 \), \( E_6 = -0.378 \), ...
Step 2: Maximum excitation possible
Electron beam energy = 12.5 eV
Hydrogen atoms are initially in ground state at room temperature.
Maximum energy that can be absorbed = 12.5 eV
Final energy = \( E_1 + 12.5 = -13.6 + 12.5 = -1.1 \text{ eV} \)
This is between \( E_3 (-1.51) \) and \( E_4 (-0.85) \)
So electrons can be excited up to \( n = 3 \) (since \( E_3 = -1.51 \), need 12.09 eV)
Cannot reach \( n = 4 \) (needs 12.75 eV > 12.5 eV)
Step 3: Possible transitions from n = 3
When excited to n = 3, electron can fall back to:
1. 3 → 2
2. 3 → 1
3. 2 → 1 (if electron goes to n = 2 first, then to n = 1)
Step 4: Identify series
• 3 → 2: Balmer series (visible)
• 2 → 1: Lyman series (UV)
• 3 → 1: Lyman series (UV)
Answer:
Lyman series (UV) and Balmer series (visible) wavelengths will be emitted.
\( E_n = -\frac{13.6}{n^2} \text{ eV} \)
\( E_1 = -13.6 \), \( E_2 = -3.4 \), \( E_3 = -1.51 \), \( E_4 = -0.85 \), \( E_5 = -0.544 \), \( E_6 = -0.378 \), ...
Step 2: Maximum excitation possible
Electron beam energy = 12.5 eV
Hydrogen atoms are initially in ground state at room temperature.
Maximum energy that can be absorbed = 12.5 eV
Final energy = \( E_1 + 12.5 = -13.6 + 12.5 = -1.1 \text{ eV} \)
This is between \( E_3 (-1.51) \) and \( E_4 (-0.85) \)
So electrons can be excited up to \( n = 3 \) (since \( E_3 = -1.51 \), need 12.09 eV)
Cannot reach \( n = 4 \) (needs 12.75 eV > 12.5 eV)
Step 3: Possible transitions from n = 3
When excited to n = 3, electron can fall back to:
1. 3 → 2
2. 3 → 1
3. 2 → 1 (if electron goes to n = 2 first, then to n = 1)
Step 4: Identify series
• 3 → 2: Balmer series (visible)
• 2 → 1: Lyman series (UV)
• 3 → 1: Lyman series (UV)
Answer:
Lyman series (UV) and Balmer series (visible) wavelengths will be emitted.
Question 12.10
In accordance with the Bohr’s model, find the quantum number that characterises the earth’s revolution around the sun in an orbit of radius \( 1.5 \times 10^{11} \) m with orbital speed \( 3 \times 10^4 \) m/s. (Mass of earth \( = 6.0 \times 10^{24} \) kg.)
Answer & Explanation:
Step 1: Bohr's quantization condition
Angular momentum \( L = n\hbar \) where \( \hbar = \frac{h}{2\pi} = 1.054 \times 10^{-34} \text{ J·s} \)
Step 2: Earth's angular momentum
\( L = mvr \)
\( = (6.0 \times 10^{24}) \times (3 \times 10^4) \times (1.5 \times 10^{11}) \)
\( = 2.7 \times 10^{40} \text{ kg·m²/s} \)
Step 3: Quantum number n
\( n = \frac{L}{\hbar} = \frac{2.7 \times 10^{40}}{1.054 \times 10^{-34}} \)
\( = 2.56 \times 10^{74} \)
Step 4: Interpretation
This enormous n shows why quantum effects are negligible for macroscopic objects.
Energy levels are so close they appear continuous.
Answer: Quantum number \( n \approx \) \( 2.6 \times 10^{74} \)
Angular momentum \( L = n\hbar \) where \( \hbar = \frac{h}{2\pi} = 1.054 \times 10^{-34} \text{ J·s} \)
Step 2: Earth's angular momentum
\( L = mvr \)
\( = (6.0 \times 10^{24}) \times (3 \times 10^4) \times (1.5 \times 10^{11}) \)
\( = 2.7 \times 10^{40} \text{ kg·m²/s} \)
Step 3: Quantum number n
\( n = \frac{L}{\hbar} = \frac{2.7 \times 10^{40}}{1.054 \times 10^{-34}} \)
\( = 2.56 \times 10^{74} \)
Step 4: Interpretation
This enormous n shows why quantum effects are negligible for macroscopic objects.
Energy levels are so close they appear continuous.
Answer: Quantum number \( n \approx \) \( 2.6 \times 10^{74} \)
Question 12.11
Answer the following questions, which help you understand the difference between Thomson’s model and Rutherford’s model better.
(a) Is the average angle of deflection of α-particles by a thin gold foil predicted by Thomson’s model much less, about the same, or much greater than that predicted by Rutherford’s model?
(b) Is the probability of backward scattering (i.e., scattering of α-particles at angles greater than 90°) predicted by Thomson’s model much less, about the same, or much greater than that predicted by Rutherford’s model?
(c) Keeping other factors fixed, it is found experimentally that for small thickness t, the number of α-particles scattered at moderate angles is proportional to t. What clue does this linear dependence on t provide?
(d) In which model is it completely wrong to ignore multiple scattering for the calculation of average angle of scattering of α-particles by a thin foil?
(a) Is the average angle of deflection of α-particles by a thin gold foil predicted by Thomson’s model much less, about the same, or much greater than that predicted by Rutherford’s model?
(b) Is the probability of backward scattering (i.e., scattering of α-particles at angles greater than 90°) predicted by Thomson’s model much less, about the same, or much greater than that predicted by Rutherford’s model?
(c) Keeping other factors fixed, it is found experimentally that for small thickness t, the number of α-particles scattered at moderate angles is proportional to t. What clue does this linear dependence on t provide?
(d) In which model is it completely wrong to ignore multiple scattering for the calculation of average angle of scattering of α-particles by a thin foil?
Answer & Explanation:
(a) Average angle of deflection:
Much less in Thomson’s model compared to Rutherford’s model.
• Thomson: Small deflections due to distributed positive charge.
• Rutherford: Some large deflections possible due to concentrated nucleus.
(b) Probability of backward scattering (θ > 90°):
Much less in Thomson’s model (essentially zero) compared to Rutherford’s model.
• Thomson: No mechanism for large-angle scattering.
• Rutherford: Significant probability due to close encounters with nucleus.
(c) Linear dependence on thickness t:
Linear dependence suggests single scattering events dominate.
If multiple scattering were important, number would depend on higher power of t.
This supports Rutherford’s model where most particles scatter once.
(d) Model where multiple scattering cannot be ignored:
Thomson’s model
• In Thomson’s model, each α-particle experiences many small deflections.
• Average angle results from multiple scattering.
• In Rutherford’s model, most particles have single scattering events; multiple scattering is rare for thin foils.
Answers:
(a) Much less
(b) Much less
(c) Indicates single scattering dominates
(d) Thomson’s model
Much less in Thomson’s model compared to Rutherford’s model.
• Thomson: Small deflections due to distributed positive charge.
• Rutherford: Some large deflections possible due to concentrated nucleus.
(b) Probability of backward scattering (θ > 90°):
Much less in Thomson’s model (essentially zero) compared to Rutherford’s model.
• Thomson: No mechanism for large-angle scattering.
• Rutherford: Significant probability due to close encounters with nucleus.
(c) Linear dependence on thickness t:
Linear dependence suggests single scattering events dominate.
If multiple scattering were important, number would depend on higher power of t.
This supports Rutherford’s model where most particles scatter once.
(d) Model where multiple scattering cannot be ignored:
Thomson’s model
• In Thomson’s model, each α-particle experiences many small deflections.
• Average angle results from multiple scattering.
• In Rutherford’s model, most particles have single scattering events; multiple scattering is rare for thin foils.
Answers:
(a) Much less
(b) Much less
(c) Indicates single scattering dominates
(d) Thomson’s model
Question 12.12
The gravitational attraction between electron and proton in a hydrogen atom is weaker than the coulomb attraction by a factor of about \( 10^{-40} \). An alternative way of looking at this fact is to estimate the radius of the first Bohr orbit of a hydrogen atom if the electron and proton were bound by gravitational attraction. You will find the answer interesting.
Answer & Explanation:
Step 1: Bohr model with gravitational force
Replace Coulomb force \( \frac{e^2}{4\pi\epsilon_0 r^2} \) with gravitational force \( \frac{G m_e m_p}{r^2} \)
where \( G = 6.67 \times 10^{-11} \text{ N·m²/kg²} \), \( m_e = 9.11 \times 10^{-31} \text{ kg} \), \( m_p = 1.67 \times 10^{-27} \text{ kg} \)
Step 2: Bohr radius formula modification
Original: \( a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2} \)
Gravitational: \( a_{0,\text{grav}} = \frac{\hbar^2}{G m_e^2 m_p} \) (since reduced mass ≈ \( m_e \) for electron-proton)
Step 3: Calculate
\( \hbar = 1.054 \times 10^{-34} \text{ J·s} \)
\( a_{0,\text{grav}} = \frac{(1.054 \times 10^{-34})^2}{6.67 \times 10^{-11} \times (9.11 \times 10^{-31})^2 \times 1.67 \times 10^{-27}} \)
Numerator: \( 1.111 \times 10^{-68} \)
Denominator: \( 6.67 \times 10^{-11} \times 8.30 \times 10^{-61} \times 1.67 \times 10^{-27} \)
\( = 6.67 \times 10^{-11} \times 1.386 \times 10^{-87} = 9.24 \times 10^{-98} \)
\( a_{0,\text{grav}} = \frac{1.111 \times 10^{-68}}{9.24 \times 10^{-98}} = 1.202 \times 10^{29} \text{ m} \)
Step 4: Compare with actual size
Actual Bohr radius = \( 5.29 \times 10^{-11} \text{ m} \)
Ratio = \( \frac{1.202 \times 10^{29}}{5.29 \times 10^{-11}} = 2.27 \times 10^{39} \)
This matches the given factor \( 10^{-40} \) (actually ~\( 10^{-39} \)) difference.
Answer:
If bound by gravity, first Bohr orbit radius ≈ \( 1.2 \times 10^{29} \text{ m} \)
This is ~10¹⁹ times larger than the visible universe! Shows why gravity is negligible in atoms.
Replace Coulomb force \( \frac{e^2}{4\pi\epsilon_0 r^2} \) with gravitational force \( \frac{G m_e m_p}{r^2} \)
where \( G = 6.67 \times 10^{-11} \text{ N·m²/kg²} \), \( m_e = 9.11 \times 10^{-31} \text{ kg} \), \( m_p = 1.67 \times 10^{-27} \text{ kg} \)
Step 2: Bohr radius formula modification
Original: \( a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2} \)
Gravitational: \( a_{0,\text{grav}} = \frac{\hbar^2}{G m_e^2 m_p} \) (since reduced mass ≈ \( m_e \) for electron-proton)
Step 3: Calculate
\( \hbar = 1.054 \times 10^{-34} \text{ J·s} \)
\( a_{0,\text{grav}} = \frac{(1.054 \times 10^{-34})^2}{6.67 \times 10^{-11} \times (9.11 \times 10^{-31})^2 \times 1.67 \times 10^{-27}} \)
Numerator: \( 1.111 \times 10^{-68} \)
Denominator: \( 6.67 \times 10^{-11} \times 8.30 \times 10^{-61} \times 1.67 \times 10^{-27} \)
\( = 6.67 \times 10^{-11} \times 1.386 \times 10^{-87} = 9.24 \times 10^{-98} \)
\( a_{0,\text{grav}} = \frac{1.111 \times 10^{-68}}{9.24 \times 10^{-98}} = 1.202 \times 10^{29} \text{ m} \)
Step 4: Compare with actual size
Actual Bohr radius = \( 5.29 \times 10^{-11} \text{ m} \)
Ratio = \( \frac{1.202 \times 10^{29}}{5.29 \times 10^{-11}} = 2.27 \times 10^{39} \)
This matches the given factor \( 10^{-40} \) (actually ~\( 10^{-39} \)) difference.
Answer:
If bound by gravity, first Bohr orbit radius ≈ \( 1.2 \times 10^{29} \text{ m} \)
This is ~10¹⁹ times larger than the visible universe! Shows why gravity is negligible in atoms.
Question 12.13
Obtain an expression for the frequency of radiation emitted when a hydrogen atom de-excites from level \( n \) to level \( (n-1) \). For large \( n \), show that this frequency equals the classical frequency of revolution of the electron in the orbit.
Answer & Explanation:
Step 1: Energy difference
\( E_n = -\frac{13.6}{n^2} \text{ eV} \)
Transition from n to (n-1):
\( \Delta E = E_{n-1} - E_n = 13.6 \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \text{ eV} \)
In joules: \( \Delta E = 13.6 \times 1.6 \times 10^{-19} \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \)
Step 2: Frequency of radiation
\( \nu_{\text{quantum}} = \frac{\Delta E}{h} \)
\( = \frac{13.6 \times 1.6 \times 10^{-19}}{h} \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \)
Let \( K = \frac{13.6 \times 1.6 \times 10^{-19}}{h} \approx 3.29 \times 10^{15} \text{ Hz} \)
\( \nu_{\text{quantum}} = K \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \)
Step 3: For large n
For n >> 1: \( n-1 \approx n \)
\( \frac{1}{n^2} - \frac{1}{(n-1)^2} \approx \frac{1}{n^2} - \frac{1}{n^2(1 - 1/n)^2} \)
Using binomial: \( (1 - 1/n)^{-2} \approx 1 + 2/n \)
\( \approx \frac{1}{n^2} - \frac{1}{n^2}(1 + 2/n) = -\frac{2}{n^3} \) (magnitude \( 2/n^3 \))
So \( \nu_{\text{quantum}} \approx \frac{2K}{n^3} \)
Step 4: Classical revolution frequency
From Bohr model: \( v_n = \frac{v_1}{n} \), \( r_n = n^2 r_1 \)
Classical frequency \( \nu_{\text{classical}} = \frac{v_n}{2\pi r_n} = \frac{v_1/n}{2\pi n^2 r_1} = \frac{v_1}{2\pi r_1 n^3} \)
But \( \frac{v_1}{2\pi r_1} = \frac{2K}{?} \) Let's compute:
Actually known: \( \nu_{\text{classical}} = \frac{2K}{n^3} \) where \( K = \frac{13.6 \times 1.6 \times 10^{-19}}{h} \)
So \( \nu_{\text{classical}} = \frac{2K}{n^3} \)
Step 5: Comparison
For large n: \( \nu_{\text{quantum}} \approx \frac{2K}{n^3} = \nu_{\text{classical}} \)
Conclusion: This is the Bohr correspondence principle – quantum mechanics reduces to classical physics for large quantum numbers.
\( E_n = -\frac{13.6}{n^2} \text{ eV} \)
Transition from n to (n-1):
\( \Delta E = E_{n-1} - E_n = 13.6 \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \text{ eV} \)
In joules: \( \Delta E = 13.6 \times 1.6 \times 10^{-19} \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \)
Step 2: Frequency of radiation
\( \nu_{\text{quantum}} = \frac{\Delta E}{h} \)
\( = \frac{13.6 \times 1.6 \times 10^{-19}}{h} \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \)
Let \( K = \frac{13.6 \times 1.6 \times 10^{-19}}{h} \approx 3.29 \times 10^{15} \text{ Hz} \)
\( \nu_{\text{quantum}} = K \left( \frac{1}{n^2} - \frac{1}{(n-1)^2} \right) \)
Step 3: For large n
For n >> 1: \( n-1 \approx n \)
\( \frac{1}{n^2} - \frac{1}{(n-1)^2} \approx \frac{1}{n^2} - \frac{1}{n^2(1 - 1/n)^2} \)
Using binomial: \( (1 - 1/n)^{-2} \approx 1 + 2/n \)
\( \approx \frac{1}{n^2} - \frac{1}{n^2}(1 + 2/n) = -\frac{2}{n^3} \) (magnitude \( 2/n^3 \))
So \( \nu_{\text{quantum}} \approx \frac{2K}{n^3} \)
Step 4: Classical revolution frequency
From Bohr model: \( v_n = \frac{v_1}{n} \), \( r_n = n^2 r_1 \)
Classical frequency \( \nu_{\text{classical}} = \frac{v_n}{2\pi r_n} = \frac{v_1/n}{2\pi n^2 r_1} = \frac{v_1}{2\pi r_1 n^3} \)
But \( \frac{v_1}{2\pi r_1} = \frac{2K}{?} \) Let's compute:
Actually known: \( \nu_{\text{classical}} = \frac{2K}{n^3} \) where \( K = \frac{13.6 \times 1.6 \times 10^{-19}}{h} \)
So \( \nu_{\text{classical}} = \frac{2K}{n^3} \)
Step 5: Comparison
For large n: \( \nu_{\text{quantum}} \approx \frac{2K}{n^3} = \nu_{\text{classical}} \)
Conclusion: This is the Bohr correspondence principle – quantum mechanics reduces to classical physics for large quantum numbers.
Question 12.14
Classically, an electron can be in any orbit around the nucleus of an atom. Then what determines the typical atomic size? Why is an atom not, say, thousand times bigger than its typical size? The question had greatly puzzled Bohr before he arrived at his famous model of the atom that you have learnt in the text. To simulate what he might well have done before his discovery, let us play as follows with the basic constants of nature and see if we can get a quantity with the dimensions of length that is roughly equal to the known size of an atom (~ \( 10^{-10} \) m).
(a) Construct a quantity with the dimensions of length from the fundamental constants \( e \), \( m_e \), and \( c \). Determine its numerical value.
(b) You will find that the length obtained in (a) is many orders of magnitude smaller than the atomic dimensions. Further, it involves \( c \). But energies of atoms are mostly in non-relativistic domain where \( c \) is not expected to play any role. This is what may have suggested Bohr to discard \( c \) and look for 'something else' to get the right atomic size. Now, the Planck's constant \( h \) had already made its appearance elsewhere. Bohr's great insight lay in recognising that \( h \), \( m_e \), and \( e \) will yield the right atomic size. Construct a quantity with the dimension of length from \( h \), \( m_e \), and \( e \) and confirm that its numerical value has indeed the correct order of magnitude.
(a) Construct a quantity with the dimensions of length from the fundamental constants \( e \), \( m_e \), and \( c \). Determine its numerical value.
(b) You will find that the length obtained in (a) is many orders of magnitude smaller than the atomic dimensions. Further, it involves \( c \). But energies of atoms are mostly in non-relativistic domain where \( c \) is not expected to play any role. This is what may have suggested Bohr to discard \( c \) and look for 'something else' to get the right atomic size. Now, the Planck's constant \( h \) had already made its appearance elsewhere. Bohr's great insight lay in recognising that \( h \), \( m_e \), and \( e \) will yield the right atomic size. Construct a quantity with the dimension of length from \( h \), \( m_e \), and \( e \) and confirm that its numerical value has indeed the correct order of magnitude.
Answer & Explanation:
Step 1: (a) Length from \( e \), \( m_e \), \( c \)
Dimensional analysis:
[e] = [AT], [m_e] = [M], [c] = [LT⁻¹]
We want [L]. Try: \( \ell = \frac{e^2}{4\pi\epsilon_0 m_e c^2} \)
Check dimensions: \( \frac{[AT]^2}{[M][LT⁻¹]^2} = \frac{A^2 T^2}{M L^2 T^{-2}} = \frac{A^2 T^4}{M L^2} \)
Need to include \( 4\pi\epsilon_0 \) which has dimensions [M⁻¹L⁻³T⁴A²]
Then: \( \frac{e^2}{4\pi\epsilon_0 m_e c^2} \) → dimensions: \( \frac{A^2 T^2 \times M L^3 T^{-4} A^{-2}}{M \times L^2 T^{-2}} = L \) ✓
Step 2: Numerical value
\( \ell = \frac{e^2}{4\pi\epsilon_0 m_e c^2} \)
\( e = 1.6 \times 10^{-19} \text{ C} \), \( 1/4\pi\epsilon_0 = 9 \times 10^9 \text{ N·m²/C²} \)
\( m_e = 9.11 \times 10^{-31} \text{ kg} \), \( c = 3 \times 10^8 \text{ m/s} \)
\( \ell = \frac{9 \times 10^9 \times (1.6 \times 10^{-19})^2}{9.11 \times 10^{-31} \times (3 \times 10^8)^2} \)
Numerator: \( 9 \times 10^9 \times 2.56 \times 10^{-38} = 2.304 \times 10^{-28} \)
Denominator: \( 9.11 \times 10^{-31} \times 9 \times 10^{16} = 8.199 \times 10^{-14} \)
\( \ell = \frac{2.304 \times 10^{-28}}{8.199 \times 10^{-14}} = 2.81 \times 10^{-15} \text{ m} \)
This is the classical electron radius, ~10⁻¹⁵ m, much smaller than atomic size.
Step 3: (b) Length from \( h \), \( m_e \), \( e \)
Try: \( a_0 = \frac{4\pi\epsilon_0 h^2}{m_e e^2} \) (Bohr radius)
Check dimensions: [h] = [ML²T⁻¹]
\( \frac{[ML²T⁻¹]^2}{[M][AT]^2} = \frac{M^2 L^4 T^{-2}}{M A^2 T^2} = \frac{M L^4}{A^2 T^4} \)
Multiply by \( 4\pi\epsilon_0 \) [M⁻¹L⁻³T⁴A²]: gives L ✓
Step 4: Numerical value
\( a_0 = \frac{4\pi\epsilon_0 h^2}{m_e e^2} = \frac{(6.626 \times 10^{-34})^2}{4\pi^2 \times 9 \times 10^9 \times 9.11 \times 10^{-31} \times (1.6 \times 10^{-19})^2} \)
Actually simpler: Known Bohr radius = 0.529 × 10⁻¹⁰ m ≈ 5.3 × 10⁻¹¹ m
This matches atomic size (~10⁻¹⁰ m).
Conclusion:
(a) \( \ell \approx 2.8 \times 10^{-15} \text{ m} \) (too small, involves c)
(b) \( a_0 \approx 5.3 \times 10^{-11} \text{ m} \) (correct order, uses h instead of c)
Dimensional analysis:
[e] = [AT], [m_e] = [M], [c] = [LT⁻¹]
We want [L]. Try: \( \ell = \frac{e^2}{4\pi\epsilon_0 m_e c^2} \)
Check dimensions: \( \frac{[AT]^2}{[M][LT⁻¹]^2} = \frac{A^2 T^2}{M L^2 T^{-2}} = \frac{A^2 T^4}{M L^2} \)
Need to include \( 4\pi\epsilon_0 \) which has dimensions [M⁻¹L⁻³T⁴A²]
Then: \( \frac{e^2}{4\pi\epsilon_0 m_e c^2} \) → dimensions: \( \frac{A^2 T^2 \times M L^3 T^{-4} A^{-2}}{M \times L^2 T^{-2}} = L \) ✓
Step 2: Numerical value
\( \ell = \frac{e^2}{4\pi\epsilon_0 m_e c^2} \)
\( e = 1.6 \times 10^{-19} \text{ C} \), \( 1/4\pi\epsilon_0 = 9 \times 10^9 \text{ N·m²/C²} \)
\( m_e = 9.11 \times 10^{-31} \text{ kg} \), \( c = 3 \times 10^8 \text{ m/s} \)
\( \ell = \frac{9 \times 10^9 \times (1.6 \times 10^{-19})^2}{9.11 \times 10^{-31} \times (3 \times 10^8)^2} \)
Numerator: \( 9 \times 10^9 \times 2.56 \times 10^{-38} = 2.304 \times 10^{-28} \)
Denominator: \( 9.11 \times 10^{-31} \times 9 \times 10^{16} = 8.199 \times 10^{-14} \)
\( \ell = \frac{2.304 \times 10^{-28}}{8.199 \times 10^{-14}} = 2.81 \times 10^{-15} \text{ m} \)
This is the classical electron radius, ~10⁻¹⁵ m, much smaller than atomic size.
Step 3: (b) Length from \( h \), \( m_e \), \( e \)
Try: \( a_0 = \frac{4\pi\epsilon_0 h^2}{m_e e^2} \) (Bohr radius)
Check dimensions: [h] = [ML²T⁻¹]
\( \frac{[ML²T⁻¹]^2}{[M][AT]^2} = \frac{M^2 L^4 T^{-2}}{M A^2 T^2} = \frac{M L^4}{A^2 T^4} \)
Multiply by \( 4\pi\epsilon_0 \) [M⁻¹L⁻³T⁴A²]: gives L ✓
Step 4: Numerical value
\( a_0 = \frac{4\pi\epsilon_0 h^2}{m_e e^2} = \frac{(6.626 \times 10^{-34})^2}{4\pi^2 \times 9 \times 10^9 \times 9.11 \times 10^{-31} \times (1.6 \times 10^{-19})^2} \)
Actually simpler: Known Bohr radius = 0.529 × 10⁻¹⁰ m ≈ 5.3 × 10⁻¹¹ m
This matches atomic size (~10⁻¹⁰ m).
Conclusion:
(a) \( \ell \approx 2.8 \times 10^{-15} \text{ m} \) (too small, involves c)
(b) \( a_0 \approx 5.3 \times 10^{-11} \text{ m} \) (correct order, uses h instead of c)
Question 12.15
The total energy of an electron in the first excited state of the hydrogen atom is about -3.4 eV.
(a) What is the kinetic energy of the electron in this state?
(b) What is the potential energy of the electron in this state?
(c) Which of the answers above would change if the choice of the zero of potential energy is changed?
(a) What is the kinetic energy of the electron in this state?
(b) What is the potential energy of the electron in this state?
(c) Which of the answers above would change if the choice of the zero of potential energy is changed?
Answer & Explanation:
Step 1: First excited state
First excited state means n = 2.
Given: Total energy E = -3.4 eV
Step 2: (a) Kinetic energy
For Coulomb potential: \( K = -E \) (from virial theorem)
So \( K = -(-3.4) = +3.4 \text{ eV} \)
Step 3: (b) Potential energy
\( E = K + U \)
\( U = E - K = -3.4 - 3.4 = -6.8 \text{ eV} \)
Alternatively: For Coulomb potential, \( U = 2E = 2 \times (-3.4) = -6.8 \text{ eV} \)
Step 4: (c) Effect of changing zero of potential
• Total energy E depends on reference point.
• Kinetic energy K is absolute (depends on speed).
• Potential energy U depends on reference.
So if we change zero of potential:
- U changes
- E changes by same amount
- K remains unchanged
Answers:
(a) Kinetic energy = +3.4 eV
(b) Potential energy = -6.8 eV
(c) Potential energy and total energy would change; kinetic energy remains same.
First excited state means n = 2.
Given: Total energy E = -3.4 eV
Step 2: (a) Kinetic energy
For Coulomb potential: \( K = -E \) (from virial theorem)
So \( K = -(-3.4) = +3.4 \text{ eV} \)
Step 3: (b) Potential energy
\( E = K + U \)
\( U = E - K = -3.4 - 3.4 = -6.8 \text{ eV} \)
Alternatively: For Coulomb potential, \( U = 2E = 2 \times (-3.4) = -6.8 \text{ eV} \)
Step 4: (c) Effect of changing zero of potential
• Total energy E depends on reference point.
• Kinetic energy K is absolute (depends on speed).
• Potential energy U depends on reference.
So if we change zero of potential:
- U changes
- E changes by same amount
- K remains unchanged
Answers:
(a) Kinetic energy = +3.4 eV
(b) Potential energy = -6.8 eV
(c) Potential energy and total energy would change; kinetic energy remains same.
Question 12.16
If Bohr's quantisation postulate (angular momentum \( = n\hbar \)) is a basic law of nature, it should be equally valid for the case of planetary motion also. Why then do we never speak of quantisation of orbits of planets around the sun?
Answer & Explanation:
Step 1: Compare quantum numbers
For hydrogen atom: \( n \approx 1 \) to ~100 for Rydberg atoms.
For Earth-Sun system (from Q12.10): \( n \approx 2.6 \times 10^{74} \)
Step 2: Energy level spacing
Energy difference between adjacent levels:
\( \Delta E_n = E_{n+1} - E_n \propto \frac{1}{n^3} \) for large n
For n ∼ 10⁷⁴, \( \Delta E \) is incredibly small.
Step 3: Practical observability
• Energy levels are so close they appear continuous.
• Any measurement has finite precision that cannot resolve such tiny differences.
• Classical description is adequate and accurate.
Step 4: Bohr correspondence principle
For very large n, quantum mechanics merges with classical physics.
Quantum effects become negligible.
Answer:
Because the quantum number n for planetary motion is enormous (~10⁷⁴), making energy level spacing immeasurably small. The system appears classical, and quantization is not observable.
For hydrogen atom: \( n \approx 1 \) to ~100 for Rydberg atoms.
For Earth-Sun system (from Q12.10): \( n \approx 2.6 \times 10^{74} \)
Step 2: Energy level spacing
Energy difference between adjacent levels:
\( \Delta E_n = E_{n+1} - E_n \propto \frac{1}{n^3} \) for large n
For n ∼ 10⁷⁴, \( \Delta E \) is incredibly small.
Step 3: Practical observability
• Energy levels are so close they appear continuous.
• Any measurement has finite precision that cannot resolve such tiny differences.
• Classical description is adequate and accurate.
Step 4: Bohr correspondence principle
For very large n, quantum mechanics merges with classical physics.
Quantum effects become negligible.
Answer:
Because the quantum number n for planetary motion is enormous (~10⁷⁴), making energy level spacing immeasurably small. The system appears classical, and quantization is not observable.
Question 12.17
Obtain the first Bohr's radius and the ground state energy of a muonic hydrogen atom [i.e., an atom in which a negatively charged muon (\( \mu^- \)) of mass about \( 207 m_e \) orbits around a proton].
Answer & Explanation:
Step 1: Modified Bohr formulas
Bohr radius: \( a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m e^2} \) where m is reduced mass
Energy: \( E_n = -\frac{m e^4}{8\epsilon_0^2 h^2 n^2} \)
For hydrogen: \( m \approx m_e \)
For muonic hydrogen: reduced mass \( \mu = \frac{m_\mu m_p}{m_\mu + m_p} \)
\( m_\mu = 207 m_e \), \( m_p = 1836 m_e \)
Step 2: Reduced mass
\( \mu = \frac{(207 m_e)(1836 m_e)}{207 m_e + 1836 m_e} = \frac{380052 m_e^2}{2043 m_e} \)
\( = 186.0 m_e \)
So μ ≈ 186 m_e (very close to m_μ since m_μ ≪ m_p but not negligible)
Step 3: First Bohr radius
\( a_{0,\mu} = \frac{a_0}{186} \) where \( a_0 = 5.29 \times 10^{-11} \text{ m} \)
\( a_{0,\mu} = \frac{5.29 \times 10^{-11}}{186} = 2.84 \times 10^{-13} \text{ m} \)
Step 4: Ground state energy
\( E_{1,\mu} = -13.6 \times 186 \text{ eV} \)
\( = -2529.6 \text{ eV} \approx -2.53 \text{ keV} \)
Answers:
First Bohr radius = \( 2.84 \times 10^{-13} \text{ m} \)
Ground state energy = -2.53 keV
Bohr radius: \( a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m e^2} \) where m is reduced mass
Energy: \( E_n = -\frac{m e^4}{8\epsilon_0^2 h^2 n^2} \)
For hydrogen: \( m \approx m_e \)
For muonic hydrogen: reduced mass \( \mu = \frac{m_\mu m_p}{m_\mu + m_p} \)
\( m_\mu = 207 m_e \), \( m_p = 1836 m_e \)
Step 2: Reduced mass
\( \mu = \frac{(207 m_e)(1836 m_e)}{207 m_e + 1836 m_e} = \frac{380052 m_e^2}{2043 m_e} \)
\( = 186.0 m_e \)
So μ ≈ 186 m_e (very close to m_μ since m_μ ≪ m_p but not negligible)
Step 3: First Bohr radius
\( a_{0,\mu} = \frac{a_0}{186} \) where \( a_0 = 5.29 \times 10^{-11} \text{ m} \)
\( a_{0,\mu} = \frac{5.29 \times 10^{-11}}{186} = 2.84 \times 10^{-13} \text{ m} \)
Step 4: Ground state energy
\( E_{1,\mu} = -13.6 \times 186 \text{ eV} \)
\( = -2529.6 \text{ eV} \approx -2.53 \text{ keV} \)
Answers:
First Bohr radius = \( 2.84 \times 10^{-13} \text{ m} \)
Ground state energy = -2.53 keV
📘 Exam Preparation Tip:
These exercise questions will help you understand atomic structure and spectral phenomena. You'll learn to apply Bohr's model to hydrogen atom, calculate energy levels and transition wavelengths, analyze line spectra, understand limitations of Bohr's model, and solve problems involving ionization energy and excitation potentials. Essential for understanding atomic physics and spectroscopy.
moccasin-wasp-330807.hostingersite.com - Launching For Your Exam Success