Fact-checked by Grok 2 weeks ago

Principal quantum number

The principal quantum number, denoted as n, is a fundamental quantum number in atomic physics that specifies the primary energy level, or shell, of an electron in an atom or ion. It takes positive integer values starting from 1 (n = 1, 2, 3, \dots), with no upper limit, and determines the average distance of the electron from the nucleus as well as the overall size of the electron's orbital. Higher values of n correspond to electrons occupying larger orbitals farther from the nucleus, resulting in increased electron energy and reduced binding to the atomic core. Introduced in Niels Bohr's 1913 model of the , the principal quantum number originally described the quantized and radius of orbits, laying the groundwork for understanding atomic spectra. In the full quantum mechanical framework developed by in 1926, n emerges as an eigenvalue of the Hamiltonian operator in the for the , where the energy levels are given by E_n = -\frac{13.6 \, \text{eV}}{n^2} for hydrogen. This quantization ensures discrete energy states, explaining phenomena such as atomic emission and absorption lines. For multi-electron atoms, n still primarily governs the shell structure, though electron-electron interactions cause deviations from the simple $1/n^2 energy dependence seen in hydrogen. Each shell defined by n can hold up to $2n^2 electrons and contains n subshells, each characterized by a secondary azimuthal quantum number l ranging from 0 to n-1. The principal quantum number thus plays a central role in electron configuration, periodic table organization, and predicting chemical properties, as electrons in higher n shells are more easily removed, influencing ionization energies and reactivity.

Fundamentals

Definition and Notation

The principal quantum number, denoted by n, is a positive integer (n = 1, 2, 3, \dots) that specifies the energy level and relative size of an atomic orbital within the quantum mechanical description of atoms. It arises as one of the quantum numbers in the solutions to the time-independent Schrödinger equation for hydrogen-like atoms, where it primarily governs the radial extent and principal energy shell of the electron's wave function. In standard notation, [n](/page/N+) takes on values beginning with , which corresponds to the of the atom, and extends theoretically without an upper bound, allowing for increasingly higher energy levels. Larger values of [n](/page/N+) indicate orbitals that are farther from the on average and encompass greater spatial volume. The designation "principal quantum number" traces its origin to early , where it served to quantize the primary or principal orbits of electrons in semi-classical atomic models. This numbering facilitates the organization of atomic states and relates directly to the quantization of energy in multi-electron atoms as well.

Role in Quantum Mechanics

In quantum mechanics, the principal quantum number n serves as the foundational descriptor for an electron's energy level and radial distribution in an , directly influencing the possible values of the other quantum numbers that specify the electron's state. Specifically, n determines the range of the l, which can take values from 0 to n-1, thereby defining the subshell types (s for l=0, p for l=1, etc.). This constraint on l in turn sets the possible values for the m_l, ranging from -l to +l in steps, which specifies the orbital's in space. The m_s, with values of +\frac{1}{2} or -\frac{1}{2}, remains independent of n but completes the set for distinguishing electron spins within the same spatial orbital. As the primary quantizer, n establishes the atomic shell structure, labeling shells as K (n=1), L (n=2), M (n=3), and so on, with each successive shell accommodating more subshells due to the increasing number of allowed l values. For instance, the K shell (n=1) has only one subshell (l=0), while the N shell (n=4) supports four subshells (l=0,1,2,3), enabling greater complexity in electron arrangements as n increases. This hierarchical organization underpins the filling of atomic orbitals according to the , where electrons occupy shells in order of rising n. In multi-electron atoms, the full specification of an electron's requires the quartet n, l, m_l, and m_s, ensuring compliance with the that prohibits two electrons from sharing identical quantum numbers. This complete set uniquely designates each electron's position and properties within the atom, facilitating the prediction of chemical behavior and spectral lines.

Historical Development

Bohr Model Introduction

In 1913, Niels Bohr developed a seminal atomic model for the , introducing the concept of quantized electron orbits to explain the discrete spectral lines observed in its . This model built on Ernest Rutherford's nuclear atom by incorporating elements of Planck's , proposing that electrons could occupy stable, non-radiating "stationary states" around the . Bohr's approach resolved a key in , where orbiting s would continuously radiate energy and spiral into the , by restricting motion to specific paths. A central postulate of Bohr's model was the quantization of the electron's in these states, expressed as mvr = n \hbar, where m is the , v its , r the , n a positive integer, and \hbar = h / 2\pi with h as Planck's constant. Here, n, later termed the , labels the allowed orbits with n = 1, 2, [3, \dots](/page/3_Dots), corresponding to circular paths of increasing radius and energy. This quantization condition ensured that only certain angular momenta—multiples of \hbar—were permissible, preventing continuous energy loss and stabilizing the . Bohr further postulated that radiation is emitted or absorbed only when the electron transitions between these quantized states, with the emitted photon's frequency given by the energy difference between levels. Applying the angular momentum quantization to the hydrogen atom, Bohr derived the energy of the n-th state as E_n = -\frac{13.6 \, \mathrm{eV}}{n^2}, where the negative value signifies bound states relative to the ionization limit at E = 0. This expression directly connects n to both the orbital radius, which scales as n^2, and the electron velocity, which scales as $1/n, providing a quantitative basis for the model's predictions of hydrogen's spectral series. The ground state at n=1 yields E_1 = -13.6 \, \mathrm{eV}, matching experimental ionization energies.

Quantum Mechanical Refinement

In 1916, extended Bohr's atomic model by incorporating elliptical orbits to account for the observed in spectral lines. This refinement introduced a subsidiary , now known as the l, which parameterized the orbit's eccentricity and , while preserving the n as the key determinant of the semi-major axis and overall energy quantization. Sommerfeld's approach retained Bohr's integer values for n but allowed for relativistic corrections, enhancing the model's explanatory power for atomic spectra without altering n's fundamental status. The transition to in the mid-1920s further refined the principal quantum number. In 1925, formulated , a non-commutative algebraic framework for quantum phenomena, in which n emerges naturally as the index labeling the discrete energy eigenvalues for stationary states in the , generalizing Bohr's ad hoc integer to a rigorous spectral label derived from matrix diagonalization. Complementing this, Erwin Schrödinger's 1926 wave mechanics solved the time-independent for the , yielding wavefunctions where n denotes the associated with energy levels, appearing as an integer eigenvalue that quantizes the radial extent and energy, thus embedding n within a continuous wave description of electron probability distributions. Discussions at the 1927 Solvay Conference on Electrons and Photons, attended by luminaries including Bohr, Heisenberg, Schrödinger, and Einstein, played a pivotal role in consolidating the principal quantum number's place in non-relativistic quantum theory. Participants debated the interpretive foundations of the new mechanics, affirming n's centrality in describing discrete energy states and transitions, thereby bridging old quantum ideas with the matrix and wave formulations and establishing n as an invariant feature of the emerging quantum paradigm.

Theoretical Derivation

Semiclassical Approach

The Bohr-Sommerfeld quantization rule provides a semiclassical framework for deriving the principal quantum number n by imposing quantization conditions on classical periodic orbits, serving as an intermediate step between and full . This rule generalizes Bohr's original angular momentum quantization to more general motions using action-angle variables. For a system with periodic motion, the action over one complete cycle must satisfy \oint \mathbf{p} \cdot d\mathbf{q} = n h, where \mathbf{p} is the canonical momentum, d\mathbf{q} is the infinitesimal displacement in configuration space, n is a positive integer known as the principal quantum number, and h is Planck's constant. This condition ensures that the phase space area enclosed by the orbit is an integer multiple of h, quantizing the allowed states and introducing discreteness into the otherwise continuous classical trajectories. In its application to the hydrogen atom, the Bohr-Sommerfeld rule builds directly on Bohr's simpler circular orbit model. For circular orbits, the quantization of angular momentum yields m v r = n \hbar, where m is the electron mass, v is its speed, r is the orbital radius, and \hbar = h / 2\pi. Solving the classical equations of motion for the Coulomb potential V(r) = -e^2 / r (in atomic units) together with this condition gives discrete orbital radii r_n \propto n^2, with the constant of proportionality being the Bohr radius a_0 = \hbar^2 / (m e^2). Thus, r_n = n^2 a_0. Sommerfeld extended this to elliptical orbits by introducing separate quantization for radial and angular actions, J_r = n_r h and J_\phi = l h, where n_r and l are radial and azimuthal quantum numbers, respectively, leading to the principal quantum number n = n_r + l + 1. This derivation reproduces the discrete shell structure of atomic orbitals, with larger n corresponding to more distant, higher-energy orbits. Despite its successes, the semiclassical approach has inherent limitations, particularly for small values of n. It provides accurate energy levels for the hydrogen atom across all n due to the special properties of the $1/r potential, but the underlying orbital descriptions break down for n=1, where the ground state (l=0) has no classical analog—the electron cannot classically orbit without angular momentum without spiraling into the nucleus. The method neglects zero-point energy effects from quantum fluctuations, which in full quantum mechanics enforce a non-zero minimum kinetic energy and prevent collapse, rendering the classical turning points invalid near r=0 for low n and low l. This approximation improves for large n, where de Broglie wavelengths are small compared to orbital scales, allowing classical-like behavior.

Wave Mechanical Derivation

The wave mechanical derivation of the principal quantum number originates from solving the time-independent for the , which describes the stationary states of the in the potential. The equation is given by -\frac{\hbar^2}{2\mu} \nabla^2 \psi + V \psi = E \psi, where \psi(\mathbf{r}) is the wave function, \mu is the of the electron-proton system, E is the energy eigenvalue, and V(r) = -\frac{e^2}{r} is the attractive potential (in cgs units). This governs the bound states with E < 0, and its solutions yield the quantized energy levels indexed by the principal quantum number n. To solve this, the wave function is separated in spherical coordinates (r, \theta, \phi), assuming \psi(r, \theta, \phi) = R(r) \Theta(\theta) \Phi(\phi), due to the spherical symmetry of the potential. Substituting into the Schrödinger equation and dividing by R \Theta \Phi leads to three ordinary differential equations: one for the azimuthal part \Phi(\phi), which yields the magnetic quantum number m, one for the polar part \Theta(\theta), which combines with \Phi to form spherical harmonics Y_{l m}(\theta, \phi) introducing the orbital quantum number l, and the radial equation for R(r). The separation constants enforce l(l+1) in the centrifugal term. The radial equation takes the form \frac{d}{dr} \left( r^2 \frac{dR}{dr} \right) + \left[ \frac{2\mu r^2}{\hbar^2} (E - V(r)) - l(l+1) \right] R = 0, or, with the substitution u(r) = r R(r), -\frac{\hbar^2}{2\mu} \frac{d^2 u}{dr^2} + \left[ V(r) + \frac{\hbar^2 l(l+1)}{2\mu r^2} \right] u = E u. For the Coulomb potential, a change of variables \rho = \frac{2}{\ a_0 n} r (where a_0 relates to the ) transforms this into the associated Laguerre differential equation. The solutions involve associated Laguerre polynomials L_{n_r}^{2l+1}(\rho), where n_r is a secondary index. The principal quantum number n emerges from the boundary conditions requiring the wave function to remain finite at r = 0 and normalizable as r \to \infty. Near r = 0, the solution behaves as R(r) \sim r^l to cancel the centrifugal singularity, while at large r, an exponential decay e^{-\rho/2} ensures square-integrability for bound states. The power series solution for the radial function terminates only if the coefficient of the recursive relation vanishes, imposing n_r = n - l - 1 where n = 1, 2, 3, \dots and n > l. This termination condition quantizes the allowed states, with n serving as the principal quantum number that indexes the eigenvalues.

Physical Properties

Energy Levels

In the quantum mechanical treatment of the hydrogen atom, the bound-state energy levels depend exclusively on the principal quantum number n, which takes positive integer values starting from 1. The energy E_n for the electron in a hydrogen atom (nuclear charge Z = 1) is given by E_n = -\frac{13.6 \, \text{eV}}{n^2}, where the negative sign indicates bound states relative to the ionization threshold at E = 0. This expression arises from the exact solution of the time-independent Schrödinger equation for the Coulomb potential and holds independently of the azimuthal quantum number l and the magnetic quantum number m_l, leading to degeneracy in those quantum numbers for each n. The ground state corresponds to n = 1, with E_1 = -13.6 \, \text{eV}, while higher n values yield less negative energies, approaching zero as n \to \infty. For hydrogen-like ions, consisting of a nucleus of charge Z e and a single electron, the energy levels scale with the square of the nuclear charge due to the strengthened Coulomb attraction in the potential V(r) = -\frac{Z e^2}{4 \pi \epsilon_0 r}. The generalized formula becomes E_n = -\frac{13.6 \, \text{eV} \, Z^2}{n^2}, again independent of l and m_l, with n = 1 as the ground state. For example, in singly ionized helium (Z = 2), the ground-state energy is -54.4 \, \text{eV}, four times deeper than in hydrogen. This scaling applies to all one-electron systems, such as He^+, Li^{2+}, and higher-Z ions. While the non-relativistic formula provides the dominant dependence on n, relativistic effects introduce fine-structure corrections that modify the levels weakly, with shifts scaling inversely with powers of n beyond $1/n^2 (such as $1/n^3 for the leading term). These adjustments, arising from the Dirac equation's treatment of electron spin and relativistic , partially lift the l-degeneracy but retain an overall primary dependence on n.

Orbital Characteristics

The principal quantum number n fundamentally governs the spatial extent of atomic orbitals in hydrogen-like atoms. The average radial distance of the electron from the nucleus, given by the expectation value \langle r \rangle, scales as \langle r \rangle \propto n^2 a_0, where a_0 is the (approximately 0.529 ). This quadratic dependence means that orbitals with higher n form larger electron shells, with the size increasing dramatically—for instance, the average radius for n=2 is roughly four times that for n=1. A key structural feature tied to n is the number of radial nodes, which are spherical surfaces where the radial probability density vanishes. The number of these nodes is exactly n - l - 1, with l being the (ranging from 0 to n-1). Radial nodes divide the orbital into regions of alternating probability, affecting how the is distributed along the ; for example, the 3p orbital (n=3, l=1) has one radial node, creating a more complex probability profile than the node-free 1s orbital./Quantum_Mechanics/09._The_Hydrogen_Atom/Radial_Nodes) Higher values of n result in more diffuse orbitals, where the electron probability is spread over a greater volume, leading to less concentrated density near the nucleus overall. For s-states (l=0), the radial wavefunction remains non-zero at the origin, enabling penetration to the nucleus regardless of n, though the increased diffuseness for larger n shifts much of the probability to greater distances./Book%3A_University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/08%3A_Atomic_Structure/8.02%3A_The_Hydrogen_Atom)

Applications and Implications

In Atomic Spectra

The principal quantum number n governs the discrete energy levels in the , enabling the prediction and interpretation of its atomic and spectra through transitions between these levels. When an jumps from a higher level n_2 to a lower level n_1 (with n_2 > n_1), it emits a with equal to the difference between the levels, producing sharp spectral lines; occurs in the reverse process. These transitions, where \Delta n \neq 0, follow the empirical for the wavenumber: \frac{1}{\lambda} = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where \lambda is the wavelength and R is the Rydberg constant for hydrogen (R \approx 1.097 \times 10^7 \, \mathrm{m}^{-1})./08%3A_Central_Potentials/8.04%3A_Rydberg_Formula) This formula, originally empirical, was theoretically justified by Niels Bohr in 1913 as arising from quantized stationary states labeled by n. The formula organizes the hydrogen spectrum into distinct series, each corresponding to a fixed lower level n_1 and varying higher levels n_2 > n_1. The (n_1 = 2) produces visible lines, such as the red H\alpha line at 656.3 nm (n_2 = 3 \to 2) and blue H\beta at 486.1 nm (n_2 = 4 \to 2), first empirically described by Johann Balmer in 1885 using measurements of 's visible emission lines. The (n_1 = 1), observed by Theodore Lyman from 1906 to 1914, lies in the region, with the strongest line at 121.6 nm (n_2 = 2 \to 1). The Paschen series (n_1 = 3) appears in the , starting at 1875 nm (n_2 = 4 \to 3). Additional series like Brackett (n_1 = 4) and Pfund (n_1 = 5) extend further into the . These series converge as n_2 increases, reflecting the increasing density of energy levels with higher n./08%3A_Central_Potentials/8.04%3A_Rydberg_Formula) As n_2 \to \infty, the energy difference approaches the from the lower level, yielding the series limit where $1/\lambda = R / n_1^2; beyond this, the becomes a corresponding to , with no discrete lines./Quantum_Mechanics/09._The_Hydrogen_Atom/Bohr's_Hydrogen_Atom) For the , this limit is at 91.2 nm, marking the onset of the ultraviolet absorption. Bohr's model predicted these limits theoretically, linking them to the as n \to \infty. Experimental verification of n-based predictions came from early spectroscopic observations, where hydrogen's discharge tube emissions matched the Rydberg formula's line positions precisely. Balmer's 1885 analysis of visible lines, later extended by Rydberg's 1888 generalization, aligned with Bohr's 1913 quantum interpretation, confirming the role of n values in producing the observed discrete spectrum without assumptions. Modern precision measurements, such as those using diffraction gratings, continue to validate these predictions to high accuracy./08%3A_Central_Potentials/8.04%3A_Rydberg_Formula)

In Multi-Electron Systems

In multi-electron atoms, the principal quantum number n continues to define the primary energy shells, but electron-electron interactions significantly modify the simple hydrogen-like ordering. The governs the sequential filling of orbitals, starting with the lowest energy levels, where orbitals are grouped by increasing n and, within each shell, by l. For example, the ground-state configuration of carbon (Z=6) is $1s^2 2s^2 2p^2, filling the n=1 shell completely before occupying n=2. This principle, formalized in the early quantum era, ensures that electrons occupy subshells in order of rising n + l values, with ties resolved by lower n first, as per Madelung's empirical ordering rule. The maximum electron capacity of each shell is given by $2n^2, arising from the degeneracy of orbitals: for a given n, there are n possible l values (0 to n-1), each subshell holds $2(2l+1) electrons due to magnetic and spin quantum numbers, summing to $2n^2 total states per Pauli exclusion./02:_Atomic_Structure/2.02:_The_Schrodinger_equation_particle_in_a_box_and_atomic_wavefunctions/2.2.03:_Aufbau_Principle) A key modification in multi-electron systems is the screening effect, where inner-shell electrons shield outer electrons from the full nuclear charge Z, reducing the Z_{\text{eff}} experienced by electrons in higher n shells. This shielding lowers the of outer orbitals relative to hydrogen-like expectations, compressing levels and influencing filling order. For an in shell n, Z_{\text{eff}} = Z - \sigma, where \sigma is the screening constant accounting for contributions from other electrons; electrons in the same shell screen less effectively (about 0.35) than those in inner shells (up to 0.85 or 1.00). provide a practical approximation for \sigma, grouping electrons into l units and assigning shielding factors based on their principal quantum numbers relative to the electron of interest—for instance, all electrons in shells with n-1 contribute 0.85 to \sigma for s or p electrons. This effective charge decrease stabilizes higher n shells less than expected, altering inter-shell spacings. Despite these guidelines, exceptions to the Aufbau filling order occur in transition metals, where the proximity in between ns and (n-1)d orbitals—due to partial screening and relativistic effects—favors configurations with half-filled or fully filled subshells for enhanced . Chromium (Z=24) adopts [\ce{Ar}] 4s^1 3d^5 instead of the expected [\ce{Ar}] 4s^2 3d^4, as the half-filled $3d^5 (lower n=3) provides greater exchange and despite the higher principal shell for $4s (n=4). Similarly, (Z=29) has [\ce{Ar}] 4s^1 3d^{10} over [\ce{Ar}] 4s^2 3d^9, prioritizing the closed $3d^{10} subshell for minimized unpaired electrons and increased . These anomalies highlight how n-dependent radial distributions influence orbital and overall in multi-electron atoms.

References

  1. [1]
    Quantum Numbers and Electron Configurations
    The principal quantum number (n) describes the size of the orbital. Orbitals for which n = 2 are larger than those for which n = 1, for example. Because they ...
  2. [2]
    Quantum Numbers and Electronic Structure
    The principal quantum number may have values as follows: n = 1, 2, 3, 4, ….. The azimuthal quantum number (also called subsidiary or secondary), l. Each ...
  3. [3]
    30.8 Quantum Numbers and Rules – College Physics
    We define n to be the principal quantum number that labels the basic states of a system. The lowest-energy state has n = 1 , the first excited state has
  4. [4]
    Quantum numbers for hydrogen atom - HyperPhysics
    The hydrogen atom solution to the Schrodinger equation produces three quantum numbers which can be seen to arise naturally from geometrical constraints on the ...
  5. [5]
    Hydrogen Atom - UT Physics
    Since we are in three dimensions, there are three quantum numbers, the principal quantum number n, an eigenvalue associated with the operator for angular ...
  6. [6]
  7. [7]
    [PDF] Chapter 10: Hydrogen Atom, Schrodinger Eq, Spherical Coordinates
    The new integer index n is known as the principal quantum number. Using the principal quantum number, it follows that the eigenenergies of the hydrogen atom are.
  8. [8]
    [PDF] Niels Bohr - Nobel Lecture
    Besides the principal quantum integer n, the stationary states are further characterized by a subordinate quantum integer s, which can be negative as well as ...
  9. [9]
    [PDF] CHAPTER 7: The Hydrogen Atom - SIBOR
    Analysis of the Schrödinger wave equation in three dimensions introduces three quantum numbers that quantize the energy. ▫. A quantum state is degenerate when ...
  10. [10]
    Quantum Numbers for Atoms - Chemistry LibreTexts
    Aug 14, 2024 · The principal quantum number, n , describes the energy of an electron and the most probable distance of the electron from the nucleus. In other ...
  11. [11]
    Quantum Numbers and Rules | Physics - Lumen Learning
    Quantum numbers are used to express the allowed values of quantized entities. The principal quantum number n labels the basic states of a system and is given by ...
  12. [12]
    Principal Quantum Number - an overview | ScienceDirect Topics
    The principal quantum number is defined as the quantum number \( n \) that determines a particle's energy level, taking on values of 1, 2, 3, and so on. AI ...<|control11|><|separator|>
  13. [13]
    Niels Bohr's First 1913 Paper: Still Relevant, Still ... - AIP Publishing
    Nov 1, 2018 · Bohr's classic first 1913 paper on the hydrogen atom and to clarify what in that paper is right and what is wrong (as well as what is weird).The need for h · Stationary states · Quantum jumps · The correspondence principle
  14. [14]
    [PDF] 1913 On the Constitution of Atoms and Molecules
    This paper is an attempt to show that the application of the above ideas to Rutherford's atom-model affords a basis for a theory of the constitution of atoms.
  15. [15]
    The hydrogen atom
    En = -13.6 eV/n2. Here n is called the principle quantum number. The values En are the possible value for the total electron energy (kinetic and potential ...
  16. [16]
    Hydrogen Atom - Galileo
    The energy E=−1/n2, the levels are labeled nl, n being the principal quantum number and the traditional notation for angular momentum l is given at the bottom ...
  17. [17]
    [PDF] {How Sommerfeld extended Bohr's model of the atom (1913–1916)}
    Jan 30, 2014 · Sommerfeld anticipated the introduction of a third quantum number n for a third phase integral concerning the z coordinate which would ...Missing: principal | Show results with:principal
  18. [18]
    [PDF] THE BIRTH OF QUANTUM MECHANICS Werner Heisenberg ...
    plays the role of the principal quantum number n which had already made its appearance in Bohr's 1913 theory of the Balmer spectrum. Four non-hydrogenic ...
  19. [19]
    Quantum Numbers (Erwin Schrödinger)
    ... Schrödinger in 1926 ... The three coordinates that come from Schrödinger's wave equations are the principal (n), angular (l), and magnetic (m) quantum numbers.
  20. [20]
  21. [21]
    [PDF] 3. Quantisation as an eigenvalue problem; by E. Schrödinger
    With this variation problem we substitute the quantum conditions. First of all, we will take for H the Hamiltonian function of the Keplerian motion and show ...
  22. [22]
    None
    ### Summary of Principal Quantum Number Derivation from Radial Schrödinger Equation for Hydrogen
  23. [23]
    [PDF] Schrödinger's original quantum–mechanical solution for hydrogen
    Jul 4, 2021 · In this work, we show how the Laplace method can be used to solve for the quantum–mechanical energy eigenfunc- tions of the hydrogen atom, ...<|separator|>
  24. [24]
    Energy Levels of Hydrogen Atom - Richard Fitzpatrick
    The energy eigenvalues are quantized, and can only take the values $\displaystyle E = \frac{E_0}{n^{\,2}},$ where $\displaystyle E_0 = - \frac{m_e\, Missing: principal | Show results with:principal
  25. [25]
    Quantisierung als Eigenwertproblem - Wiley Online Library
    Quantisierung als Eigenwertproblem. E. Schrödinger,. E. Schrödinger ... Download PDF. back. Additional links. About Wiley Online Library. Privacy Policy ...
  26. [26]
    Principle Quantum Number - Chemistry 301
    where Z is the nuclear charge. For hydrogen (Z=1), the energy levels start at n=1 with a value of -2.18 x 10-18J.Missing: principal | Show results with:principal
  27. [27]
    [PDF] Quantum Physics III Chapter 2: Hydrogen Fine Structure
    Apr 2, 2022 · For each fixed value of ℓ, the states have increasing N as we move up in energy. The number N is the number of nodes in the solution of the ...
  28. [28]
    D.83 Expectation powers of r for hydrogen
    This note derives the expectation values of the powers of $r$ for the hydrogen energy eigenfunctions $\psi_{nlm}$.
  29. [29]
    I. On the constitution of atoms and molecules - Taylor & Francis Online
    On the constitution of atoms and molecules. N. Bohr Dr. phil. Copenhagen. Pages 1-25 | Published online: 08 Apr 2009.
  30. [30]
    [PDF] Johann Jacob Balmer (1825-1898) - UB
    Using measurements by H. W. Vogel and by Huggins of the ultraviolet lines of the hydrogen spectrum I have tried to derive a formula which will represent the ...
  31. [31]
    The Spectral Lines of Hydrogen | Spectroscopy Online
    The late 19th century empirical formulas by Balmer and Rydberg on the spectral lines of hydrogen provided crucial confirmation of the strength of Bohr's early ...
  32. [32]
    Atomic Shielding Constants | Phys. Rev.
    Simple rules are set up giving approximate analytic atomic wave functions for all the atoms, in any stage of ionization.Missing: John | Show results with:John
  33. [33]
    Exceptions to Electron Configuration Rules - The Physics Classroom
    The two most commonly mentioned exceptions to the Aufbau Principle are the elements chromium ( 24 Cr) and copper ( 29 Cu).