Docsity
Docsity

Prepare for your exams
Prepare for your exams

Study with the several resources on Docsity


Earn points to download
Earn points to download

Earn points by helping other students or get them with a premium plan


Guidelines and tips
Guidelines and tips

Understanding the Electronic Structure of Atoms and the Pauli Exclusion Principle, Schemes and Mind Maps of Physics

How the solutions of the Schrödinger equation for atoms can be characterized by four quantum numbers: n, β, m, and ms. It introduces the Pauli exclusion principle, which states that two identical fermions cannot occupy the same state, and discusses how it applies to the electronic structure of atoms, including the ground state of helium and lithium. The document also touches upon the filling of subshells in the periodic table and the combination of angular momenta.

Typology: Schemes and Mind Maps

2021/2022

Uploaded on 09/12/2022

virgyn67
virgyn67 🇺🇸

4.3

(12)

6 documents

1 / 8

Toggle sidebar

Partial preview of the text

Download Understanding the Electronic Structure of Atoms and the Pauli Exclusion Principle and more Schemes and Mind Maps Physics in PDF only on Docsity! - - 116 Chapter 7. Atomic Physics Notes: • Most of the material in this chapter is taken from Thornton and Rex, Chapter 8. 7.1 The Pauli Exclusion Principle We saw in the previous chapter that the hydrogen atom could be precisely understood by considering the Schrödinger equation, including the electrostatic potential energy that accounts for the interaction between the nucleus and the electron. We then discovered that the family of stationary states, which are the solutions of the Schrödinger equation can be completely characterized by three quantum numbers n, , and m . However, we also found that the state of the electron is further characterized by its intrinsic spin, which acts in a way akin to the angular momentum in that it provides an extra magnetic moment and is quantized similarly. But comparisons between the Ŝ and L̂ operators should probably not be pushed too far, as for the electron the spin is half-integer with s = 1 2 independent of any other parameter, while 0 ≤  ≤ n −1 and always an integer. As a result, the states of the hydrogen can only be fully specified by combining the previous three quantum numbers for the solutions of the Schrödinger equation and the magnetic spin quantum number ms , forming the n,,m,ms( ) foursome. As we will now discover there is another very important aspect of the spin that is absolutely essential for understanding the structure of many-electron atoms. Although it would be in principle possible to solve the Schrödinger equation for more complicated atoms, the presence of several interaction terms between the different electrons makes the problem analytically intractable and basically impossible to solve. The exact determination of the stationary states, their energy, angular momentum, etc. must then be accomplished using computers. It is, however, possible to qualitatively understand the structure of many-electron atoms using the results obtain for the study of the hydrogen atom and another fundamental principle that we owe to the Austrian physicist Wolfgang Pauli (1900-1958). The so-called Pauli exclusion principle, which stems from Pauli’s efforts to explain the structure of the periodic table, can be stated as follows Two identical fermions cannot occupy the same state. Remember that fermions have half-integer spins; the electron with s = 1 2 is therefore one. As Pauli initial enunciation of his principle in 1925 happened as he was studying the atomic structure, it can then stated more specifically for atomic electrons with No two electrons in an atom can share the same set of quantum numbers n,,m,ms( ) . - - 117 The structure of atoms and that of the periodic table can be explained with this principle and the further assumption that atomic electrons tend to occupy the lowest available energy states. To see how this works, let us consider the next simplest atom after hydrogen, i.e., helium. The helium atom (He) is composed of a nucleus made of two protons and two neutrons for a total charge of +2e (a neutron has the same mass as a proton but no charge) and two electrons. As was the case for the hydrogen atom we can expect that the lowest energy state for an electron will consists of a 1s wave function with the electron spin quantum number either 1 2 or −1 2 , i.e., n,,m,ms( ) = 1,0,0,±1 2( ) . The second electron is also likely to occupy a similar state, but because of the Pauli exclusion principle, which forbids two electrons from occupying the exact same state, the quantum numbers can only be n,,m,ms( ) = 1,0,0,1 2( ) ; note the opposite sign of ms for the two electrons. We are thus left with the picture of the helium atom having its two electrons in 1s states, where one electron has its spin “up” and the other its spin “down.” The complete ground state of the helium atom is then denoted by 1s2 , where the ending superscript specifies the number of electrons in the given state (in this case, 1s). What would be the electronic structure for the next simplest atom, i.e., Lithium (Li), which contains three electrons (and a nucleus made of three protons and four neutron)? It should now be clear that the Pauli exclusion principle will forbid the lowest atomic state to be something like 1s3 , since this would imply that two electrons would have to share the same ms number (i.e., the ending superscript cannot be greater than two for a ns orbital). To minimize energy two electrons will still occupy the inner shell 1s2 , while the third one will reside on the next unoccupied orbital with favourable energy. In this case the next available lowest-energy electron state is 2s. The lithium atom ground state is therefore 1s22s1 . One would be justified to ask why couldn’t the ground configuration be 1s22p1 instead? Indeed, this appears to be supported by the fact that our solution for the hydrogen atom asserted that s and p orbitals for a given n number have the same energy En . The answer lies with the consideration of the precise shapes of the different orbitals. An s orbital is said to be more penetrating than a p orbital. That is, the radial wave function for  = 0 is more concentrated closer to the nucleus than that for  = 1 . This implies that an electron on a 2p orbital in the lithium atom is more likely to have the nuclear charge of +3e screened by the two electrons on the 1s orbitals and “feel” an effective charge of  +e . On the other hand, an electron on the more penetrating 2s orbital is not as screened and will see more of the nuclear charge (i.e., the effective charge is greater than +e ), which results in a lower potential energy due to its stronger interaction with the nucleus. The 1s22s1 state is thus of lower energy than the 1s22p1 state and the correct choice for the ground state of lithium. Just as the orbitals are designated by letters depending on the values of the  quantum number, e.g., s, p, d, f, etc., shells are associated to the different values of the principal quantum number n . More precisely, levels of n = 1, 2, 3, 4,… are given the capital letters K, L, M, N, … The aforementioned n orbitals are then called subshells. It - - 120 their electrons easily removed render them quite chemically reactive. They have a valence of +2 and are good electrical conductors. Groups 3 to 12 – Transition Metals (or Transition Elements) The three rows where the 3d, 4d, and 5d subshells are being filled (i.e., elements 21-30, 39-48, and 72-80) form the transition metals group. Several of these atoms (e.g., iron (Fe), cobalt (Co), and nickel (Ni)) have strong magnetic moments due to the presence of unpaired electrons in the d subshell. These electrons will see their spins aligned therefore producing the ferromagnetic properties of the elements. The rare earths elements consisting of the lanthanides (elements 58-71) and actinides (elements 90-103) can also be included in this group. This is because these elements have unpaired electrons in the f subshells (for n = 4 and n = 5 , respectively) leading also to large magnetic moments. Group 17 - Halogens All elements of this group have five electrons in their outer p subshell, and therefore have a valence of −1. This characteristic renders them very chemically reactive; Fluorine (F) is the most reactive element in existence. Halogens will especially bond efficiently with alkalis, which have a valence of +1, to form compound such as NaCl. Atoms in groups 13 to 16 are composed of metals (e.g., aluminum (Al), tin (Sn), and bismuth (Bi)), non-metals (carbon (C), nitrogen (N), and oxygen (O)), and metalloids (often semiconductors) exhibiting some properties of metals and non-metals (e.g., boron (B), silicon (Si), Arsenic (As), and tellurium (Te)). 7.3 The Combination of Angular Momenta We saw that transitions metals have high magnetic moments because of the effect of unpaired electron’s spin. To understand how this comes about we must first understand how angular momenta, orbital and intrinsic spin, combine or add up to form the total angular momentum Ĵ . We will consider the simple case for the combination of the spin Ŝ and orbital angular momentum L̂ of a single electron. We first note that the two momenta add vectorially Ĵ = L̂+ Ŝ. (7.1) We remember that both the orbital and spin angular momenta are quantized such that L =  +1( ), Lz = m S = s s +1( ), Sz = ms (7.2) - - 121 where m ≤  and ms ≤ s . We therefore expect that the total angular momentum will also be quantized in a similar manner with J = j j +1( ) Jz = mj, (7.3) Fi gu re 1 – T he p er io di c ta bl e, w ith th e el ec tro ni c co nf ig ur at io n sp ec ifi ed fo r e ve ry e le m en t. - - 122 and mj ≤ j . From equations (7.2) and (7.3) we can write Jz = m +ms( ), (7.4) or mj = m +ms . It is important to realize that different values for m , ms , or mj imply a different orientation for the corresponding vectors. We therefore expect that there will be several possibilities for both the orientation of the total angular momentum Ĵ as well as its magnitude J . To get a better sense of this let us consider the case where  = 1 , m = −1, 0, and 1 , and s = 1 2 , ms = −1 2 and 1 2 . Considering equation (7.4) tells us that the following values for mj are realized mj = 3 2,1 2 m=1 ms=± 1 2   ,1 2,−1 2 m=0 ms=± 1 2   ,−1 2,− 3 2 m=−1 ms=± 1 2    . (7.5) These values for the magnetic total quantum number can be grouped as follows to find the realized values for j according to equation (7.3) j = 3 2 , mj = 3 2 , 1 2 ,− 1 2 ,− 3 2 j = 1 2 , mj = 1 2 ,− 1 2 . (7.6) This result can be generalized to any pair of angular momenta of any kind (i.e., any mixture of orbital, spin, or “intermediate” total angular momenta) with Ĵ = Ĵ1 + Ĵ2 j1 − j2 ≤ j ≤ j1 + j2 mj ≤ j, (7.7) where successive values for J differ by 1. It can easily be verified that equations (7.6) are verified when Ĵ1 = L̂1 and Ĵ2 = Ŝ1 . Let us now come back to the case of a transition metal atom and see if we can better understand its high magnetic moment relative to elements of other groups. For example, we consider the case of titanium (Ti), which has the Ar[ ]3d2 4s2 electronic configuration ( Ar[ ] means that the inner core of titanium corresponds to the filled electronic configuration of argon, which is 1s22s22p6 3s2 3p6 ). The d subshell of titanium is thus
Docsity logo



Copyright © 2024 Ladybird Srl - Via Leonardo da Vinci 16, 10126, Torino, Italy - VAT 10816460017 - All rights reserved