Collective Effects in Condensed Matter Physics 1st Edition Vladimir V. Kiselev - The Ebook in PDF Format With All Chapters Is Ready For Download
Collective Effects in Condensed Matter Physics 1st Edition Vladimir V. Kiselev - The Ebook in PDF Format With All Chapters Is Ready For Download
com
https://ebookgate.com/product/collective-effects-in-
condensed-matter-physics-1st-edition-vladimir-v-kiselev/
OR CLICK BUTTON
DOWLOAD EBOOK
https://ebookgate.com/product/encyclopedic-dictionary-of-condensed-
matter-physics-1st-ed-edition-charles-p/
ebookgate.com
https://ebookgate.com/product/isotope-effects-in-solid-state-
physics-1st-edition-vladimir-g-plekhanov-eds/
ebookgate.com
https://ebookgate.com/product/a-quantum-approach-to-condensed-matter-
physics-1st-edition-philip-l-taylor/
ebookgate.com
https://ebookgate.com/product/magnetism-in-condensed-matter-1st-
edition-stephen-blundell/
ebookgate.com
Forces growth and form in soft condensed matter at the
interface between physics and biology 1st Edition A.
Skjeltorp
https://ebookgate.com/product/forces-growth-and-form-in-soft-
condensed-matter-at-the-interface-between-physics-and-biology-1st-
edition-a-skjeltorp/
ebookgate.com
https://ebookgate.com/product/condensed-matter-field-theory-2ed-
edition-altland-a/
ebookgate.com
https://ebookgate.com/product/advances-in-amorphous-semiconductors-
advances-in-condensed-matter-science-1st-edition-jai-singh-editor/
ebookgate.com
https://ebookgate.com/product/aperiodic-structures-in-condensed-
matter-fundamentals-and-applications-1st-edition-enrique-macia-barber/
ebookgate.com
https://ebookgate.com/product/polymer-thin-films-series-in-soft-
condensed-matter-first-edition-ophelia-k-c-tsui/
ebookgate.com
Vladimir V. Kiselev
Collective Effects in Condensed Matter Physics
De Gruyter Studies in
Mathematical Physics
|
Editor in Chief
Michael Efroimsky, Bethesda, Maryland, USA
Leonard Gamberg, Reading, Pennsylvania, USA
Dmitry Gitman, São Paulo, Brazil
Alexander Lazarian, Madison, Wisconsin, USA
Boris Smirnov, Moscow, Russia
Volume 44
Vladimir V. Kiselev
Collective Effects
in Condensed
Matter Physics
|
Mathematics Subject Classification 2010
82D, 81V80, 74B
Author
Dr. Sci. (Phys.-Math.) Vladimir V. Kiselev
Russian Academy of Sciences
M. N. Mikheev Institute of Metal Physics
Sophia Kovalevskaya str., 18
620108 Yekaterinburg
[email protected]
ISBN 978-3-11-058509-4
e-ISBN (PDF) 978-3-11-058618-3
e-ISBN (EPUB) 978-3-11-058513-1
ISSN 2194-3532
www.degruyter.com
Foreword
To date, numerous phenomena and processes observable on a macroscopic scale have
been studied in condensed matter. However, to explain them properly, quantum con-
cepts about the microstructure of the medium are required. The interconnection and
mutual complementation of different fields of knowledge is one of the noteworthy fea-
tures of the current state of science. All the monographs, on the one hand, are devoted
to special problems of condensed matter physics. They give a deep and complete expo-
sition, for example, of the physics of semiconductors, superconductivity, or quantum
optics intended for experienced specialists in a narrow area of solid state physics. The
purpose of other publications is to draw readers’ attention to the fundamental prob-
lems of quantum macrophysics and its numerous technological applications. The lack
of discussion about constructive aspects of the theory, arising difficulties, and possible
ways to overcome them is a disadvantage of most works on this theme. To successfully
resolve the problems of quantum macrophysics in practice, it is necessary to overview
a huge amount of material as quickly and thoroughly as possible, which requires great
effort. There is an acute shortcoming in the scientific literature and methodical tech-
niques on quantum macrophysics for students, graduate students, and researchers.
A distinctive feature of this book is that it covers not only selected issues deemed
to be a necessary part of the education of a scientific researcher, but it also discusses
the ways of a gradual approach to solving the problems considered, i.e., what usually
remains behind the scenes. The discussion of specific issues is brought about through
complementary levels of theoretical description, quantum, semiclassical, and phe-
nomenological approaches. This makes it possible to more fully address the multi-
faceted nature of macroscopic quantum effects and, ultimately, come closer to the
problems of the current state of science within this field.
This book is based on a course of lectures the author has been delivering for a
number of years to students of the Theoretical Physics and Applied Mathematics De-
partment of the Faculty of Physics and Technology at Ural Federal University. The
specifics of the existing education system is that third year students (future engineers
or scientists) who seek a bachelor’s degree should have the knowledge of condensed
matter physics. It is this connecting role that is played by the course “Quantum Macro-
physics”. The exercises are chosen, not only for educational purposes, but also in or-
der to supplement the main material.
This book was first published in the scientific and educational series, Condensed
Matter Physics, initiated and edited by V.V. Ustinov, a Member of the Academy of Sci-
ence and a director of the Institute of Metal Physics of the Ural Branch of the Russian
Academy of Sciences. The present edition is intentionally supplemented with a new
chapter: Dislocations and Martensitic Transitions. It includes analysis of dislocation
systems, solutions of the strongly nonlinear and nonlocal Peierls model for dislocation
https://doi.org/10.1515/9783110586183-201
VI | Foreword
cores, and the discussion of the soliton mechanism of anomalous acoustic emission
near the martensitic phase transition point. The author is very grateful to the adminis-
tration of the Institute of Metal Physics for creating favorable conditions for work and
to D.V. Dolgikh and A.A. Raskovalov for technical assistance in preparing the book.
Contents
Foreword | V
Prefactory Notes | X
3 Superconductivity | 154
3.1 The Basic Physical Properties of Superconductors | 154
3.2 The Qualitative Features of the Microscopic Theory | 159
3.3 The Second Order Correction to the Energy of a Two Electron System
Due to Electron-Phonon Interaction | 162
3.4 Cooper Pairs | 166
3.5 The Bardeen–Cooper–Schrieffer Theory (Qualitative Results) | 170
3.6 The Ginzburg–Landau Theory – The London Penetration Depth | 174
3.7 Quantization of a Magnetic Flux | 177
3.8 The Microscopic Nature of Two Types of Superconductors – Vortex
Lattices and Superconducting Magnets | 178
3.9 Possible Physical Mechanisms of High Temperature
Conductivity | 182
3.10 High Temperature Superconductors | 185
Exercises | 345
Bibliography | 357
Index | 361
Prefactory Notes
1. The Concept of Quasiparticles as Collective Perturbations
of a Medium
https://doi.org/10.1515/9783110586183-202
Prefactory Notes | XI
m r2̇⃗s
εs = + U(r ⃗s ) .
2
In the frame of classical physics, we can trace in detail the motion of each of the
material points of an ideal gas by solving the system of equations of motion with given
initial conditions. However, this process involves serious computational difficulties.
Fortunately, it is not absolutely necessary to follow this process. When the number of
particles in the system is large, experimentally observable parameters can be ascer-
tained, not through intricate paths of the individual particles, but as a result of their
collective action. Collective properties do not depend on the initial conditions of the
particle motion, and are well described in terms of averages and probabilities.
The course of classical statistical physics shows that if particles of a thermally
equilibrium ideal gas, being in an external potential field, are in a state with the en-
ergy ε s , their average number can be determined by the Boltzmann distribution:
μ − εs
f s = exp ( ) , (1)
kB
where k B is the Boltzmann constant and T and ε s are the gas temperature and the en-
ergy of a single particle, respectively. The parameter μ is called the chemical potential
and is determined by the normalization condition. The latter expresses the conserva-
tion of the total number of particles N in the system:
∑ fs = N .
s
Here we note an important point. The definition of an ideal gas neglects the poten-
tial energy of interaction between its particles. At the same time, the equilibrium dis-
tribution of particles of an ideal gas in energy (the Boltzmann distribution) is caused
by the interaction of the gas particles with the thermostat walls and their collisions
with each other.
Prefactory Notes | XIII
The above conception has impelled scientists to use statistical methods for de-
scribing quantum mechanical systems of identical particles, which has led to the
emergence of classical statistical physics. When the initial wave function of a sys-
tem consisting of N identical particles is known, the system’s evolution is possible to
watch by solving the Schrödinger equation:
∂ ̂⃗ 1 , q̂⃗ 1 , . . . , p
̂ (p ̂⃗ N , q̂⃗ N ) Ψ ,
iℎ Ψ=H
∂t
where t is time, Ψ is the wave function of the system, and p ̂⃗ i , q̂⃗ i is the coordinate and
momentum operators of the i-th particle. Unlike classical physics, we are currently
unable to follow the trajectories of the particles. However, in principle, by solving the
Schrödinger equation with the initial wave function, we can detect the particle system
at any time.
Nonetheless, the above procedure is rarely feasible, even when the number of the
particles is small. This is due to not knowing the initial conditions with the desired
accuracy. In the case of a large number of the particles, any attempts to estimate the
multiparticle wave function are doomed to failure as, for example, 1 cm3 of a typical
metal (Cu, Ag, Al) contains 1023 electrons.
Instead, it is useful to turn to the statistical properties of an ensemble of identical
particle systems differing from each other only in the initial conditions of motion of
these particles. In doing so, we should employ the ergodic hypothesis. The statistics of
the ensemble is assumed to allow the mean time behavior of the given particle system
to be predicted as the latter is in a state of thermal equilibrium with its surroundings
(we replace averaging over time by the averaging over the ensemble).
Consider a quantum mechanical system consisting of N weakly interacting parti-
cles enclosed in a volume V. The physical nature of the particles is of no interest to
us, even if the particles were quanta of electromagnetic field (photons). Suppose the
density of the particles are small enough, with good approximation, we can regard the
system energy as the sum of the energies of the individual particles.
Under the conditions formulated above, the laws of quantum statistical physics
give only two of the most probable distributions of particle in energies.
1. For N identical particles with half integer spin, the average number of the par-
ticles with the energy ε s (and a given spin projection if ε s is spin independent) at a
given temperature T is determined by the Fermi–Dirac distribution:
1
fs = ε −μ
. (2)
exp ( ksB T ) +1
Just as in classical physics, the chemical potential μ can be found from the nor-
malization condition:
∑ fs = N .
s
XIV | Prefactory Notes
It is important that the summation is over all admissible quantum mechanical states
of the system, including spin ones. In particular, when spin is 1/2, the energy ε s does
not depend on the spin and each summand is present in the sum twice.
2. An ideal gas of N particles with integer spin. At thermal equilibrium, the aver-
age number of bosons with the energy ε s (and a given spin projection, if ε s does not
depend on the spin) is determined by the distribution of the Bose–Einstein:
1
fs = ε −μ
. (3)
exp ( ksB T ) −1
The latter differs from the Fermi–Dirac distribution only in the sign before unity in the
denominator.
The total number of particles in an ideal Bose gas is given by:
∑ fs = N .
s
The above formula is responsible for the chemical potential μ. The summation is also
over all admissible quantum mechanical states, including spin ones.
Through the course of statistical physics, we can see that the chemical potential μ
formally introduced by us is related to the second law of thermodynamics. Namely,
for systems with a variable number of particles, the second law of thermodynamics is
written in the following form:
Here, we can see the correspondence principle between more general quantum
physics and classical physics limited in their ability.
It is known that the classical Boltzmann distribution is often not acceptable to
describe quasiparticles at real temperatures and energies. Ideal gases of quasiparti-
cles obey quantum statistics. This causes the exhibiting of their unusual macroscopic
properties (from the point of view of experience of everyday life). It should be empha-
sized that the case refers to systems of micro- or quasiparticles whose mutual interac-
tion can be neglected.
Our next goal is to understand why the approximation of ideal Fermi and Bose
gases describes observable macroscopic properties of condensed matters surprisingly
well, consisting of a large number of strongly interacting electrons and ions.
At the first sections of the book we tell about the electronic properties of metals
and semiconductors. And only after that we examine the dynamics of a crystal lattice.
In our view, such a sequence of exposition allows one to explain the fruitfulness of the
concept of quasiparticles and the necessity of utilizing special methods to set forth pe-
riodic structures. In doing so, we will resort to relatively simple examples. A distinctive
feature of this book is that the author not only selected issues required for the educa-
tion of a scientist, but also considered gradual approaches for solving the problems at
hand. In other words, the author discloses what usually remains “behind the scenes.”
Most of the problems of quantum macrophysics begins with the application of quali-
tative methods being the most attractive and beautiful trait of this science. It is about
building simple models, which correctly take into account the main interactions and
peculiarities of the problems, but at the same time, despite their simplicity, provide
exact solutions. The author cannot help touching various estimation techniques, the
study of the limiting cases when the smallness of a parameter can be used, and the
extraction of corollaries from the symmetry properties of a medium, etc. This book dis-
cusses the particular issues through mutually complementary levels of the theoretical
description such as quantum, quasiclassical, and phenomenological approaches. All
of the foregoing topics make it possible to investigate macroscopic quantum effects
more thoroughly, pictorially illustrating the universality and equivalency of the theo-
retical approaches (in the general area of applicability), and, ultimately, come closer
to exploring the problems of the current state of modern science.
1 Electrons and Holes in Metals and Semiconductors
1.1 Electrons in Crystalline Solids: The Formulation of a
Simplified Single Particle Model
In condensed matters, the outer shells of atoms overlap. Therefore, the valence elec-
trons of the atoms can move from one atom to another. Such an overlapping explains
the high bond energy of crystalline solids and their specific mechanical properties.
The inner electron shells of atoms in crystals overlap slightly; they are almost the same
as isolated atoms.
It is important to note that inner shell electrons, and even almost free (valence)
electrons, interact strongly with each other and with ions of the crystal lattice. The
potential energy of the interactions is in the order of the kinetic energy of the electrons.
In this case, the theory of crystalline solids, at first glance, seems quite impossible to
construct. However, there is currently quite a rigorous description of most phenomena
in crystals. The reasons for this success can be explained as follows:
I. It turns out that the theoretical description of a system of strongly interacting
electrons and ions may often be reduced to the analysis of a simpler problem. Namely,
the analysis of the behavior of noninteracting quasiparticles as being fermions in a
weak external field needs to be accomplished. The latter represents the average field
of the lattice and the other electrons.
There are two main reasons why the strong interaction of the valence electrons
between both themselves and the lattice ions leads to the resultant effect, describable
by the weak potential:
a) In the permitted zone, due to their mobility in a crystal, the valence elec-
trons diminish the total potential by acting on an individual electron. Firstly, they
can screen the positive charge of ions, thereby reducing the total potential. Secondly,
the screening of the electron-electron interaction occurs. Due to Coulomb repulsion,
every electron pushes the rest of the other electrons away from itself. Therefore, a re-
gion that contains the positive uncompensated charge of fixed lattice ions surrounds
such an electron. The cloud of the positive charge moves along with it. Every electron
“sees” not only the other electrons but the time dependent polarization cloud created
by them. As a result, the effective interaction between the electrons is dynamically
screened.
Ultimately, it appears that the valence electrons in a crystal have nothing to do
with free electrons in a vacuum. They turn into quasiparticles: fermions. The latter
differ in mass from electrons in a vacuum and almost do not interact with each other
at distances of interatomic order.
b) The second reason is more compelling and associated with the Pauli principle.
The interaction between electrons and ions is strongest at small distances. However,
the Pauli principle forbids valence electrons from being near the ions, as this area
https://doi.org/10.1515/9783110586183-001
2 | 1 Electrons and Holes in Metals and Semiconductors
is already occupied by electrons of the ion core. The Pauli principle also limits the
number of collisions between conduction electrons. As a consequence, the conduction
electrons rarely suffer scattering by each other.
Given the importance of these reasons, we shall return later to discuss them.
II. Analysis of the behavior of electrons in a crystal is simplified by employing a
single particle model. If one picks up a quasiparticle weakly interacting with its sur-
roundings, an electron from the whole system, then describes the effect of all the other
particles on it through an effective potential energy, the initial many particle problem
is reduced to a system of a single particle Schrödinger equation:
ℎ2
− ⃗ = εΨ ,
∆Ψ + U(r)Ψ
2m
where m is the mass of the quasiparticle, correlated to the electron, Ψ being a station-
ary wave function of the quasiparticle, and ε being the energy.
It turns out that most observable macroscopic properties and the most important
phenomena in solids can be explained by the rational choice of the effective poten-
tial U(r)⃗ for a single electron in the Schrödinger equation. Many implications follow
from the fact that the average field of the ions and other electrons possess the symme-
try properties of the crystal lattice; periodicity, in particular.
So, the periodic potential is assumed to model an impact of the periodic field of the
crystal ions on the selected electron, as well as periodic effects caused by the action
of the rest of the electrons on the given electron.
It should be emphasized that perfect periodicity of U(r)⃗ is an idealization. Real
solids are never absolutely pure. They contain impurities and, in the neighborhood
of the impurity atoms, the solid is not the same as elsewhere in the crystal. The real
crystals have various types of defects. The ions, in fact, continually undergo thermal
vibrations around their equilibrium positions. All the disturbances of the periodicity
must be taken into account. For example, based on these disturbances, we can explain
why the conductivity of a metal is not infinite. However, to construct the theory, the
problem is artificially divided in two parts:
1. At the outset, we should consider a hypothetical perfect crystal with absolute pe-
riodicity.
2. Furthermore, we should investigate the effect of all possible deviations from the
periodicity, treated as small perturbations, on the properties of the crystal. This is
necessary in order to describe the real macroscopic properties of crystal bodies.
III. The periodicity of the potential U(r)⃗ is one of the reasons why the electrons ex-
perience no scattering by the lattice of a crystal. In a perfect periodic crystal, the free
motion of the electron can be described through the electronic wave function. This is
because any linear wave in any periodic structure propagates freely and faces the peri-
odic structure only in special cases. In the last case, reflection/scattering takes place.
This will be discussed in more detail later.
1.2 The Theoretical Description of the Periodic Structure of Crystals | 3
The behavior of electrons in the periodic potential can be examined not only in
metals. Most of our general conclusions also refer to dielectrics and semiconductors.
To go further in the analysis of the properties of a gas of noninteracting quasipar-
ticles, fermions in the periodic potential, we should get familiar with methods of the
theoretical description of the crystal lattices in solids.
The existence of a periodic ion lattice is of fundamental significance for modern solid-
state physics. The periodical ion lattice underlies all analytical constructions. If such a
lattice did not exist, the theory of crystals would hardly reach serious success. Because
of the lack of the periodic arrangement of ions, the theory of amorphous solids is still
in its infancy. Achievements in the theory of liquids are also poor, although these two
forms of matter have a density close to the density of crystalline solids.
To describe infinite crystal structures, a special symbolic language was created.
An ideal crystal can be built up in the space by periodically repeating identical struc-
tural units. The simplest crystals have a structural unit consisting of a single atom.
In crystals of more complex substances, the structural unit may contain a few atoms
or molecules. We will describe the crystal structure in terms of a crystal lattice unit
cell periodically recurring in space. It is known as an elementary cell, which binds a
certain group of atoms. This group of atoms is called a basis; the basis is repeated in
space and forms the crystalline structure.
Although the crystal structure is formed by the repetition of real physical objects (a
certain group of atoms or ions), for a theoretical description of the periodic structure
of a crystal, the abstract concept of a Bravais lattice is used. The latter, in principle,
may have no atoms in its nodes. The Bravais lattice concept reflects only geometry of
a regularly distributed array of crystal elements whatever these in fact are.
We give two equivalent definitions of the Bravais lattice.
A Bravais lattice (two-dimensional or three-dimensional) is an infinite periodic
structure formed by discrete mathematical points and has the exact same spatial order
and orientation when viewed from any lattice point [1]).
A Bravais lattice (two-dimensional or three-dimensional) is a set of mathematical
points with the radius vectors of the form:
Q
c1 a1
a2
P R
c2
Fig. 1.1: A two-dimensional crystal, formed by atoms located at the ver-
tices of hexagonal “honeycombs”.
where a⃗ i are any linearly independent vectors and n i are arbitrary integers. The vec-
tors a⃗ i are called basis vectors of the Bravais lattice.
To better understand the difference between a real crystal formed by physical ob-
jects, atoms, and a Bravais lattice, a two-dimensional crystal rather than a three-di-
mensional one is easier to start with. For example, the simplest two-dimensional crys-
tal is a layer of atoms adsorbed on the crystal face – a substrate.
Let us consider a two-dimensional crystal formed by atoms located at the vertices
of hexagonal “honeycombs” (Figure 1.1). This structure seems to be the same if one
looks at it from the points P and Q. However, the views from the point R and the point P
differ in a 180 degree turn. It means that, by the first definition, the vertices of the
honeycombs do not form the Bravais lattice.
On the other hand, it is easy to check that no two linearly independent vectors
connecting the atoms of the crystal, such as c⃗1 and c⃗ 2 (Figure 1.1), generate the Bra-
vais lattice. Therefore, by the second definition, the vertices of the two-dimensional
honeycombs do not form the Bravais lattice.
In this case, the Bravais lattice is generated by the vectors a⃗ 1 and a⃗ 2 not connect-
ing any real atoms. In Figure 1.1, the asterisks mark the Bravais lattice points and the
bold dots indicate real atoms. This example demonstrates the difference between the
Bravais lattice whose nodes, in general, contain abstract mathematical points and the
natural crystal, the nodes of which are always occupied with atoms or ions.
The term “a Bravais lattice” can be used not only to refer to a set of the points,
but also to denote the set of vectors, connecting any of these points with the rest. Any
vector R⃗ n of (1.1) defines a translation (shift) when the entire aggregation of the sys-
tem’s elements moves as a whole in space at a distance R n towards the vector R⃗ n . The
translation vectors of the crystal lattice bind different points of the Bravais lattice. In
practice, from the problem situation, it is always clear whether there are either points
or vectors or translations.
If, being subjected to all translations of the form:
a volume of space fills up the entire space/plane, never overlaps with itself and leaves
no voids, it is referred to as a primitive (elementary) unit cell.
1.2 The Theoretical Description of the Periodic Structure of Crystals | 5
a2
b2 Fig. 1.2: The elementary unit cell of the two-dimensional crystal shown in
b1 Figure 1.1; b⃗ 2 = 2b⃗ 1 , a⃗ 1 + a⃗ 2 = 3 b⃗ 1 , | a⃗ 1 | = | a⃗ 2 |, the angle between the
a1 vectors a⃗ 1 and a⃗ 2 is 60∘ .
In general, each primitive cell has only one point of the Bravais lattice. Therefore, if n
is the density of the mathematical points of the Bravais lattice, and V a is the volume
of the unit cell, the following formula holds:
1
V a n = 1 or Va = . (1.2)
n
Hence an important conclusion suggests itself. Due to the result (1.2) being valid for
any primitive cell, the unit cell volume is independent on the choice of the cell.
It also follows from the definition of a primitive cell, that for two primitive cells of
arbitrary shape one of them can be always cut into pieces, which after shifting by the
appropriate lattice vectors are join together into another primitive cell.
There are many possible ways of choosing a primitive cell. Choosing a primitive cell
in the shape of a parallelepiped has its drawback. The latter is that it does not reflect
the full symmetry of the Bravais lattice. We can always choose a primitive cell so that
it would possess the symmetry of the Bravais lattice. An example of such a choice is
the Wigner–Seitz cell.
The Wigner–Seitz cell, centered at some point (a mathematical point of the Bravais
lattice), is the region of space that lies closer to that point than to any of the other
lattice points.
When being subjected to translations through all lattice vectors, the Wigner–Seitz
cell fills up the entire space without any overlapping, i.e., the Wigner–Seitz cell is a
primitive cell.
The following simple method offers the construction of the Wigner–Seitz cell. Here
are a few steps to do it [6].
1. Choose any point of the Bravais lattice.
2. From the chosen center (the Bravais lattice point is centered), we draw the trans-
lation vectors to the nearest lattice nodes.
3. Draw planes perpendicular to these vectors, passing through their centers.
The resulting cell that takes the smallest volume containing the given point bounded
by the built planes is the primitive Wigner–Seitz cell.
Note that not all nearest lattice points can be used to construct the Wigner–Seitz
cell (Figure 1.4).
1.3 A Reciprocal Lattice and the First Brillouin Zone | 7
All points being equivalent, the Bravais lattice must be infinite in extent. Clearly,
actual crystals have finite dimensions. However, if they are large enough, most of the
points are so far from the surface as to be unaffected by its existence. The sample is
often treated as the finite lattice with N nodes to be the set of points:
R⃗ n = n1 a⃗ 1 + n2 a⃗ 2 + n3 a⃗ 3 ,
where 0 ≤ n1 ≤ N1 , 0 ≤ n2 ≤ N2 , 0 ≤ n3 ≤ N3 and N = N1 N2 N3 . Further in this Chap-
ter, we shall justify such a choice of the finite size of the crystal and related boundary
conditions.
K⃗ = n1 b⃗ 1 + n2 b⃗ 2 + n3 b⃗ 3 , (1.5)
The elementary (primitive) Wigner–Seitz cell for the reciprocal lattice is called the
first Brillouin zone (see Fig. 1.5–1.8). As the name implies, it can be specified as the fol-
lowing Brillouin zones, which are the elementary cells of a different kind. Higher Bril-
louin zones appear in the theory of electron energy levels in a periodic potential. The
terms “the Wigner–Seitz cell” and “the first Brillouin zone” are attributed to the same
geometric constructions. The last term is used only to refer to a cell in the reciprocal
space.
a
Fig. 1.5: A one-dimensional direct lattice
0 x (a chain of atoms).
2π
a
0 k
The first Brillouin zone Fig. 1.6: A one-dimensional reciprocal lattice.
a
0 x Fig. 1.7: A square direct lattice (atoms are located in the crystal face –
a a substrate).
ky
The first
Brillouin zone
2π
a
0 kx
2π
a Fig. 1.8: A square reciprocal lattice.
1.3 A Reciprocal Lattice and the First Brillouin Zone | 9
Atomic Planes
There is a close relationship between the vectors of the reciprocal lattice and atomic
planes of the direct lattice. A plane passing through mathematical points of a direct
Bravais lattice is called an atomic plane. A family of atomic planes of a lattice is the
set of equally spaced parallel atomic planes. We will demonstrate that the reciprocal
lattice makes it very easy to classify all possible families of the atomic planes. To do
this, we first explain the geometrical meaning of the reciprocal lattice vectors. Let vec-
tors of the form R⃗ m = a⃗ 1 m1 + a⃗ 2 m2 + a⃗ 3 m3 generate a direct Bravais lattice, m i are
integers; a⃗ 1 , a⃗ 2 , a⃗ 3 being vectors of the basic translations. K⃗ = b⃗ 1 n1 + b⃗ 2 n2 + b⃗ 3 n3 is
a fixed vector of the reciprocal lattice, where b⃗ i are vectors of the basic translations in
reciprocal space. Next, we are interested in only the direction given by the vector K.⃗
Therefore, from all vectors parallel to K,⃗ we choose only that for which the numbers
n1 , n2 , n3 contain no common divisors and give the smallest length of the vector K.⃗
The vectors { R⃗ N } define a set of points of the direct lattice, which are solutions of
the equation:
where N is an integer. For given N and K,⃗ the number m i runs a certain set of values.
From analytic geometry it is known that equation (1.7) defines a plane perpendicular
to the vector K.⃗ In our case, this equation is that of the atomic plane passing through
the nodes of the direct lattice with a normal line parallel to the vector K.⃗
Taking into account that:
̂⃗ ̂⃗
A⃗ ⋅ B⃗ = A⃗ B⃗ cos(A⃗ B)⃗ = A⃗ (PrB⃗ A) = B⃗ (PrA⃗ B) ,
where N is an integer and PrK⃗ R⃗ N is the projection of the vector R⃗ N onto the direction
of the vector K.⃗ The nearest integer to N is either (N + 1) or (N − 1). For the same K,⃗
the equation R⃗ N+1 ⋅ K⃗ = 2π(N + 1) determines the atomic plane in the direct lattice.
The atomic plane is parallel to the original plane and is at a minimum distance from
it. The equation of this plane can be written as:
(PrK⃗ R⃗ N+1 ) K⃗ = 2π(N + 1) . (1.9)
Let us find the distance between the nearest atomic planes perpendicular to the
vector K⃗ (Figure 1.9):
d = PrK⃗ R⃗ N+1 − PrK⃗ R⃗ N = 2π/ K⃗ . (1.10)
10 | 1 Electrons and Holes in Metals and Semiconductors
K
PrK R N +1
RN
R N +1
General Conclusions
1. Any vector K⃗ of the reciprocal lattice defines a family of atomic planes perpendic-
ular to it.
2. The minimum length of the vectors parallel to the vector K⃗ is the distance between
the nearest planes of the appropriate family of atomic planes (1.10).
The correspondence between the reciprocal lattice vectors and the families of atomic
planes gives a convenient way to specify the orientation of the atomic plane. The plane
orientation is given by the normal vector to the plane. It is natural to choose the re-
ciprocal lattice vector K⃗ as a normal to the atomic plane. Out of all vectors parallel to
vector K,⃗ the shortest vector should be selected. By doing so, the Miller indices of the
plane can be estimated.
Coordinates of the shortest reciprocal lattice vector, perpendicular to the family
of atomic planes of the direct lattice in the coordinate system specified by the basic
vectors of the reciprocal lattice, are called the Miller indices.
If K⃗ = b⃗ 1 n1 + b⃗ 2 n2 + b⃗ 3 n3 is the shortest vector, then n1 , n2 , n3 are the Miller
indices. Therefore, the Miller indices defined must be integers. This is because any re-
ciprocal lattice vector is a linear combination of three basic vectors taken with integer
coefficients. The Miller indices depend on the choice of the basic vectors. The set of
the indices (n1 , n2 , n3 ) can mean a single plane or a family of parallel planes.
If V(r ⃗ + R⃗ l ) = V(r)⃗ is the periodic potential for all R⃗ l of the Bravais lattice, then the
eigenstates of the Schrödinger equation
ℎ2
[− ⃗ Ψ = εΨ
∆ + V(r)]
2m
may be selected so that the wave functions should have the form of the plane wave
exp(ik⃗ ⋅ r),
⃗ multiplied by a function with periodicity of the Bravais lattice:
Here
u n k⃗ (r ⃗ + R⃗ l ) = u n k⃗ (r)⃗ (1.12)
for any R⃗ l belonging to the Bravais lattice, n is a natural number, with the energy
⃗
eigenvalues being ε = ε n (k).
The relations (1.11) and (1.12) can be written as follows:
̂ = − ℎ ∆ + V(r)⃗
2
H
2m
can be chosen in such a way that each of them should be associated with a wave vec-
tor k,⃗ and for any R⃗ l of the Bravais lattice, the equality should be fulfilled:
Let us explain this further. The theorem states that, although the potential in the
single particle Schrödinger equation is periodic, the eigenwave functions, generally
speaking, are not periodic. However, they have an interesting structure: they are sim-
ilar to the plane wave exp(ik⃗ ⋅ r)⃗ describing a free particle with the wave vector k.⃗
General Conclusions
1. Bloch’s theorem introduces a wave vector k,⃗ which is an analogue of the wave
vector of the free microparticle in the general problem of motion in a periodic
potential.
12 | 1 Electrons and Holes in Metals and Semiconductors
Energy gap
n-th band
Fig. 1.10: Schematic depiction of the permitted energy values ε n (k).
2. The periodic function u n k⃗ (r)⃗ with the index n modulates any wave exp(ik⃗ ⋅ r).
⃗ As
will be shown below, the Schrödinger equation has a set of discrete solutions for
each fixed k.⃗ The index n enumerates these states and is said to be a band number.
The name is due to the fact that for a fixed n the particle’s eigenenergies ε n (k), de-
pending on k,⃗ are densely clustered along the energy axis. In doing so, they form
a band separated from an n+1 neighboring band by the forbidden energy region;
an energy gap (Figure 1.10). Therefore, the eigenwave functions and eigenvalues
of the particle’s energies in a periodic potential have two indices. One of those is
the particle’s wave vector k,⃗ another is the number n of the energy band.
Since the equation (1.15) is satisfied identically for each function Ψ, the following op-
erator identity holds true:
T̂ ⃗l H
̂=H
̂T ̂ ⃗l . (1.16)
R R
Furthermore, we note that the result of two successive translations never depends
⃗ we have:
on the order of their application, since for any function Ψ(r),
̂ ⃗ l⃗ T ̂ ⃗l T
̂ ⃗l Ψ = T ̂ ⃗ ⃗l ⃗ l ⃗ ̂
T R R R R⃗ l⃗ Ψ = Ψ ( r + R + R ) = T R⃗ l + R⃗ l⃗ Ψ . (1.17)
Therefore, we obtain:
̂ ⃗l T
̂ ⃗l = T
̂ ⃗ l⃗ T ̂ ̂
T R R R R⃗ l⃗ = T R⃗ l + R⃗ l⃗ . (1.18)
1.4 Energy Levels of an Electron in a Periodic Potential and Bloch’s Theorem | 13
The relations (1.16) and (1.18) show that the Hamiltonian H ̂ and the operators T
̂ ⃗
R
⃗
for all Bravais lattice vectors R form a set of commuting operators. According to the
commuting operator theorem, the operators H ̂ and T̂ ⃗ have a common system of the
R
eigenfunctions:
̂ = εΨ , T
HΨ ̂ ⃗ Ψ = c(R)Ψ
⃗ . (1.19)
R
̂ ⃗ l (T
T ̂ ⃗ l (c(R⃗ l ⃗ )Ψ) = c(R⃗ l ⃗ )T
̂ ⃗ l⃗ Ψ) = T ̂ ⃗ l Ψ = c(R⃗ l ⃗ )c(R⃗ l )Ψ = c (R⃗ l + R⃗ l ⃗ ) Ψ . (1.20)
R R R R
Consequently, for the translation operator eigenvalues, the equality must be fulfilled:
⃗ ⃗
c (R⃗ l + R⃗ l ) = c(R⃗ l )c(R⃗ l ) . (1.21)
⃗
where R⃗ l , R⃗ l are arbitrary vectors belonging to the Bravais lattice.
As a translation vector, we take an arbitrary s-th basic vector of the direct Bravais
lattice. Then, we can always write c(a⃗ s ) as:
should be performed for each vector R⃗ l of the Bravais lattice. This is the Bloch theorem.
The wave function must remain bounded under arbitrary translations in the infi-
nite crystal, so the vector k⃗ of the representation (1.24) must be real.
14 | 1 Electrons and Holes in Metals and Semiconductors
Ψ n k⃗ (r ⃗ + N s a⃗ s ) = Ψ n k⃗ (r)⃗ , (1.25)
where a⃗ s is the trio of the basis vectors of the crystal’s Bravais lattice, N s are positive
integers,(N s ≫ 1), s = 1, 2, 3, N = N1 N2 N3 is the total number of elementary unit
cells in the crystal. In this section, there is no summation over twice repeated indices.
According to the Born–Karman boundary conditions, an electron reaches the
boundary of the sample without reflecting but “disappears” and, at the same time,
reenters the opposite face of the crystal.
We demonstrate that the boundary conditions (1.25) are consistent with the Bloch
theorem. We apply the Bloch theorem (1.13) to the boundary condition (1.25):
Then we have:
⃗ =1.
exp (iN s a⃗ s ⋅ k) (1.27)
exp [2πiN s x s ] = 1 .
b⃗ 1 b⃗ 2 b⃗ 3 Vb
δV k⃗ = ⋅[ × ]= . (1.29)
N1 N2 N3 N
Since V b = b⃗ 1 ⋅ (b⃗ 2 × b⃗ 3 ) is the volume of the elementary unit cell of the reciprocal
lattice, formula (1.29) means that the amounts of allowed wave vectors in one unit cell
and of cells in the crystal (N) are equal.
1.4 Energy Levels of an Electron in a Periodic Potential and Bloch’s Theorem | 15
In the direct and reciprocal lattices, the volumes of the unit cells are related by:
(2π)3
Vb = . (1.30)
Va
Therefore, formula (1.29) can be rewritten as:
V b (2π)3 (2π)3
δV k⃗ = = = , (1.31)
N Va N V
where V = NV a is the volume of the crystal.
Thus, the greater the volume of the crystal V is, the closer the allowed wave vectors
in k-space are to each other.
1. We have shown that the stationary Schrödinger equation with the periodic poten-
⃗
tial V(r):
ℎ2
− ∆Ψ + V(r)Ψ ⃗ = εΨ (1.32)
2m
has the solution:
Ψ n k⃗ (r)⃗ = u n k⃗ (r)⃗ exp (ik⃗ ⋅ r)⃗ ,
In formula (1.35), all the allowed wave vectors k⃗ reside near the average value of k:⃗
⃗
g(k⃗ ) ≈ 0 if k − k⃗ ≥ ∆k .
It is known from quantum mechanics that, when localized in k-space in a region with
a characteristic size of the order ∆k, a wave packet is localized in a normal coordinate
space in a region size of:
1
∆R ∼ . (1.36)
∆k
16 | 1 Electrons and Holes in Metals and Semiconductors
This region moves as a whole with a group velocity, which is calculated by the formula:
1 ∂ε n (k)⃗
V⃗ = . (1.37)
ℎ ∂ k⃗
The velocity (1.37) is given by the gradient ε n (k)⃗ of the dispersion relation at the point
with the radius vector k.⃗ This is because the harmonic components of the wave packet
have the wave vectors close to the vector k.⃗
We have come up with an interesting outcome. It turns out that, in the periodic
crystal, the electron has a steady state of being in which it moves with a constant ve-
locity without changing its energy. Also, we cannot regard collisions between elec-
trons and fixed ions as a mechanism for governing the change in the electron’s ve-
locity, since equation (1.33) (the solution of which is the function (1.34)) completely
takes into account the electron-lattice interaction. Such a result is incompatible with
classical ideas; it is a consequence of the wave nature of electrons.
In the periodic lattice of scattering centers, the wave propagates without damping
due to the interference of scattered waves. Since the resultant wave (1.35) correspond-
ing to the electron in the crystal does not decay, the conductivity of an ideal crystal
should be infinite. The cause of electrical resistance of metals is crystal lattice defects
or impurities or lattice ion vibrations. All of them disrupt the periodicity of the poten-
tial energy of the electron.
2. Bloch’s theorem introduces the wave vector k⃗ into theory. The quantity ℎk⃗ is of-
ten called the quasimomentum of an electron. For a free electron, the quasimomentum
and momentum are related: p⃗ = ℎk.⃗ However, this formula does not hold for an elec-
tron in a crystal: p⃗ ≠ ℎk.⃗ This is clear from the following considerations. The existence
of the connections ε = ε n (k), ⃗ p⃗ = ℎk⃗ would mean that the energy and momentum
of the electron can be measured simultaneously. In other words, the energy and mo-
mentum operators, corresponding to these values commute: [H, ̂⃗ = 0. In turn, the
̂ p]
̂ ̂
condition [H, p]⃗ = 0 would mean that the momentum of the electron is an integral
of motion. This is impossible, since the momentum is related to the invariance of the
electron Hamiltonian H ̂ with respect to arbitrary translations. Under the discontinu-
ous periodic potential, the electron Hamiltonian H ̂ does not possess full translational
invariance, despite the preservation of its invariance with respect to discrete transla-
tions, reflecting the symmetry of the crystal lattice. We have arrived at a contradiction
because our initial assumption of the connection p⃗ = ℎk⃗ is incorrect. For an electron
in a crystal this connection is p⃗ ≠ ℎk.⃗
By direct verification, we can see that, for the electron in an inhomogeneous crys-
tal field, the following stands: p⃗ ≠ ℎk.⃗ We act by the momentum operator p ̂⃗ = −iℎ∇⃗ on
the Bloch electron wave function:
Ψ n k⃗ = u n k⃗ (r)⃗ exp (ik⃗ ⋅ r)⃗ .
As a result, we get:
̂⃗ ⃗ ⃗ ⃗ ⃗ ⃗
pΨ n k⃗ = ℎ kΨ n k⃗ − iℎ exp (i k ⋅ r) ∇u n k⃗ ≠ ℎ kΨ n k⃗ .
1.4 Energy Levels of an Electron in a Periodic Potential and Bloch’s Theorem | 17
This means that the quasimomentum of the electron in a crystal has nothing in com-
mon with its momentum: p⃗ ≠ ℎk.⃗
To comprehend the meaning of the term “quasimomentum,” the reaction of Bloch
electrons to an external electromagnetic field should be considered. We will come back
to this issue later.
3. We can always choose the wave vector k⃗ of the Bloch function so that it should
lie within the first Brillouin zone. If the wave vector k⃗ does not fall into the first Bril-
louin zone, it can always be represented as the sum of the wave vector k⃗ of the first
Brillouin zone and any reciprocal lattice vector K:⃗
k⃗ = k⃗ + K⃗ ,
where K⃗ = b⃗ 1 n1 + b⃗ 2 n2 + b⃗ 3 n3 .
As an example, consider a square lattice. In Figure 1.11, the vector k⃗ lies outside
the first Brillouin zone. The vector k⃗ can be paired with the vector k⃗ being inside the
first Brillouin zone if one subtracts the reciprocal lattice vector K⃗ from k⃗ . The wave
vector of the point A at the zone boundary can be translated to the point A on the
opposite edge of the same zone by adding some vector K.⃗ Then the question arises of
whether we can think that both points A and A belong to the first zone? The answer
is that they are said to be identical, so we consider only one of them.
Suppose we have the Bloch function with the wave vector k⃗ outside the first Bril-
louin zone:
Ψ n k⃗ (r)⃗ = u n k⃗ (r)⃗ exp (ik⃗ ⋅ r)⃗ . (1.38)
We replace k⃗ by (k⃗ + K)⃗ and rewrite formula (1.38) in the following form:
Ψ n k⃗ (r)⃗ = u n k⃗ (r)⃗ exp (i k⃗ ⋅ r)⃗ = exp (ik⃗ ⋅ r)⃗ [exp (i K⃗ ⋅ r)⃗ u n k⃗ (r)]
⃗ = exp (ik⃗ ⋅ r)⃗ u n k⃗ (r)⃗ .
(1.39)
Here we have introduced the designation u n k⃗ (r)⃗ = u n k⃗ (r)⃗ exp(i K⃗ ⋅ r). ⃗
By virtue of the relation K⃗ ⋅ R⃗ = 2πm (m is an integer), the function exp(i K⃗ ⋅ r)⃗ as
well as u n k⃗ (r) is a periodic function in the Bravais lattice. So, u n k⃗ (r) is periodic too:
u n k⃗ (r ⃗ + R)⃗ = u n k⃗ (r)⃗ .
ky
π
a
K
k k'
π π kx
a a
A' K A
π
a
Fig. 1.11: The first Brillouin zone for a plane square lattice.
18 | 1 Electrons and Holes in Metals and Semiconductors
Instead of the original Bloch function with the wave vector k⃗ , we have obtained an
equivalent function with the wave vector k⃗ in the first Brillouin zone:
u k⃗ (r ⃗ + R)⃗ = u k⃗ (r)⃗ .
The operator H ̂ ⃗ in (1.40) is Hermitian: Ĥ+ = H ̂ ⃗ . The superscript “+” indicates the
k k⃗ k
Hermitian conjugate operation.
Due to the periodic boundary conditions, equation (1.40) can be solved within a
single elementary unit cell of the direct lattice. Thus, we have the eigenvalue problem
for a fixed finite volume. Any such problems have a discrete spectrum. This explains
the origin of the index n of the function ε n (k)⃗ and u n k⃗ (r)⃗ (n = 1, 2, . . . ).
With the operator H ̂ ⃗ being Hermitian, its eigenvalues ε = ε n (k)⃗ are real. The
k
eigenvalue problem (1.40) contains the wave vector k⃗ only as a parameter. So, the func-
tion ε n (k)⃗ is assumed to depend continuously on k.⃗ Recall that the wave vectors of the
type k⃗ = ∑3i=1 m i /N i b⃗ i can take different values under the Born–Karman boundary
conditions if the size of the crystal changes. In the limit of the infinite crystal, these
values involve a dense set in the k-⃗ space.
5. Since the vectors k⃗ and (k⃗ + K)⃗ are physically equivalent, the energy ε n (k)⃗ must
be a periodic function in reciprocal space:
with the periods b⃗ 1 , b⃗ 2 , b⃗ 3 of the basic translations. Recall that any reciprocal lattice
vector has the form K⃗ = n1 b⃗ 1 + n2 b⃗ 2 + n3 b⃗ 3 , where n1 , n2 , n3 are integers.
Thus, we come to the description of the energy levels of the electron through a
family of the continuous functions of ε n (k), ⃗ each of which has a periodicity of the
reciprocal lattice. These functions determine the band structure of solids.
Each function of ε n (k)⃗ is periodic and continuous in k.⃗ Consequently, it has upper
and lower limits and, therefore, all the energy levels of ε n (k)⃗ for a given n lie between
these two limits. Zones with different values of n can be separated by energy gaps but
may overlap.
1.4 Energy Levels of an Electron in a Periodic Potential and Bloch’s Theorem | 19
̂ = − ℎ ∇ + V(r)⃗ .
2
̂ ⃗
HΨ n k⃗ = ε n (k)Ψ n k⃗ , H
2m
(1.41)
̂ are real (H
Since the eigenvalues of ε n (ε∗n = ε n ) and the operator H ̂ ∗ = H),
̂ from (1.41)
we have:
̂ ∗ = ε n (k)Ψ
HΨ ⃗ ∗ .
⃗
nk ⃗ nk
Hereinafter, the symbol “∗ ” denotes the operation of complex conjugation. The Bloch
function determines the wave vector k:⃗
⃗
It is clear that, to the wave function Ψ∗ ⃗ , there corresponds the wave vector (−k).
nk
Therefore, the following chain of equalities holds true:
⃗ ∗ = ε n (−k)Ψ
̂ ∗ = ε n (k)Ψ
HΨ ⃗ ∗ .
⃗ nk ⃗ nk ⃗ nk
Consequently, the eigenvalues of the electron energy operator are invariant under the
replacement k⃗ → −k:⃗
ε n (k)⃗ = ε n (−k)⃗ .
7. Crystalline lattices of metals have considerable symmetry, owing to which ex-
tremes of the functions ε n (k)⃗ are at the boundaries of the Brillouin zone. For lattices
of special symmetry there is also an additional extremum at the center of the Brillouin
zone. We can make sure of this on particular examples.
The ground state of a system of Bloch electrons in a crystal can be constructed accord-
ing to the following principles:
1. The principle of minimum energy (electrons fill up allowed energy levels succes-
sively, starting with the lower available level).
2. The Pauli Principle.
Single-electron energy levels of ε n (k)⃗ depend on two quantum numbers n and k⃗ and
meet the condition: ε n (k)⃗ = ε n (k⃗ + K).
⃗ Wave vectors that differ by a reciprocal lattice
20 | 1 Electrons and Holes in Metals and Semiconductors
Empty bands
εg Energy gap
vector are physically equivalent and correspond to the same energy value. Therefore,
when taking each energy level into account only once, we must restrict the values k⃗ to
the first Brillouin zone. In addition, the Pauli exclusion principle reads that no more
than two electrons may occupy the same energy level of ε n (k)⃗ simultaneously. To oc-
cupy the same energy level, two electrons with quantum numbers n, k⃗ must have op-
posite spins. Consequently, this stands for different quantum states.
By virtue of the above principles, the electrons distributed in the energy levels can
form the following configurations:
1. Some bands may appear to be completely filled with electrons and the remain-
ing may appear to be empty. (Figure 1.12). The energy difference εg between the “top”
of the highest occupied band and the “bottom” of the lowest empty band is referred
to as a forbidden band or an energy gap.
A solid satisfying the condition εg ≫ k B T is an insulator. However, if εg ≈ k B T, it
is an intrinsic semiconductor.
The number of levels in the band is equal to the number of elementary cells (N)
in a crystal. Every level has two electrons. Consequently, the maximum number of
energetic states in every band is 2N. To obtain a semiconductor or an insulator, it is
necessary that the number of electrons per unit cell should be even. Otherwise, not all
quantum mechanical states will be filled up with electrons in the upper of the bands.
This is a necessary condition for semiconductors and dielectrics to be formed, how-
ever, it is not sufficient due to possible overlapping of the bands.
2. Some bands may appear to be partially filled (Figure 1.13). Such crystalline
solids are metals.
A crystal with an even number of valence electrons per unit cell should be re-
garded as having both overlapping and nonoverlapping energy zones. Each case
needs to be considered separately. If the zones overlap, we may obtain two or more
1.4 Energy Levels of an Electron in a Periodic Potential and Bloch’s Theorem | 21
Fig. 1.13: The distribution of electrons among energy levels. There are par-
tially filled zones.
partially filled ones instead of a full band, inherent in a dielectric. This results in
exhibiting the metal properties by the crystal.
The energy of the highest occupied level is called the Fermi energy εF . In three-di-
mensional crystals, the Fermi energy range can cover some zones. For every partially
filled band in the k-space there is a surface which separates occupied states from un-
occupied states. Such a surface is called the Fermi surface.
Analytically, the Fermi surface is evaluated by the equation:
ε n (k)⃗ = εF . (1.42)
The Fermi surface is not necessarily 1-connected; it can exist in the form of indi-
vidual sections, having different values of the index n in formula (1.42). In this case,
the number of such sections lying around the partially filled zones is always small.
The shape of the Fermi surface is responsible for most of the electronic properties
of metals as an electric current occurs by changing the number of occupied electron
states near the Fermi surface.
Because ε n (k)⃗ is a periodic function in reciprocal space, the solution of (1.42),
which defines the Fermi surface, has the same property. The Fermi surface is periodic
with the periods b⃗ 1 , b⃗ 2 , b⃗ 3 in the k-space.
There are different ways to depict the Fermi surface geometrically. If one draws the
Fermi surface for all admissible vectors k,⃗ a periodic surface will be produced. This
is called the repetitive band method. For clarity, one can illustrate a periodic Fermi
surface within the first Brillouin zone. In this case, each physically different energy
electron level is represented by only one point at the Fermi surface; no level will be
lost or counted twice. Such a representation of the Fermi surface is characteristic for
the so called reduced zone scheme.
Analogously, any periodic function can be represented, such as sin x, either on
the real axis (the repetitive zone scheme) or within one of the periods (the reduced
zone scheme). We will return to this subject later.
22 | 1 Electrons and Holes in Metals and Semiconductors
Let us consider the problem of the motion of a single electron in a crystal in the pres-
ence of a weak periodic potential. In this case, the electron motion is almost free. As
such, we can gain a complete understanding of the energy spectrum of the electron
and the Fermi surface through the perturbation theory. Metals of groups I-IV in the
periodic table have s and p valence electrons, and their inner atomic orbitals are com-
pletely filled up with electrons. Therefore, such metals are well suited for building the
nearly free electron model to qualitatively explain the behavior of the electrons within
them.
A metallic state is one of the most important states of matter. Chemical elements
clearly prefer to be in the metallic state: more than two-thirds of them are metals.
Therefore, in spite of being more applicable to metals than to semiconductors, it is
difficult to overestimate the significance of the nearly free electron model.
Before proceeding to the basic subject matter, we will take a little mathematical
excursus.
Periodic Function Expansion into Plane Waves for Multiple Measurement Cases
Any function possessing the periodicity property can be expanded into plane waves,
which in turn form a complete set of functions. In a one-dimensional space, a func-
tion f(x) with the above property f(x) = f(x + L) can be represented as a Fourier series:
The situation is similar in the three-dimensional case. If the function f(r)⃗ has the
same periodicity as the Bravais lattice:
The integration in (1.44) is carried out over a volume V a of each elementary cell of the
direct lattice. Strictly speaking, the choice of the cell is of no consequence because the
integrand in (1.44) is periodic. The integral of the periodic function of the elementary
cell is cell independent. In fact, any primitive cell can be split into parts. By translating
the latter by the Bravais lattice vectors, we can obtain other primitive cells. Moreover,
when translated, the periodic function exp(i K⃗ ⋅ r)f(
⃗ r)⃗ remains unchanged.
At this point it would be appropriate to prove the identity:
1
∫ d3 r ⃗ exp [i(K⃗ − K⃗ ) ⋅ r]⃗ = δ K,⃗ K⃗ . (1.45)
Va
Va
The expression under the integral sign in (1.45) has the same periodicity as the Bravais
lattice. Therefore, a shift of the integration region by an arbitrary vector d⃗ does not
change the integral. By integrating exp[i(K⃗ − K⃗ ) ⋅ r]⃗ over a mixed cell of V a , we can
⃗
arrive at an integral over the original cell of V a of exp[i(K⃗ − K⃗ ) ⋅ (r ⃗ + d)]:
⃗ .
∫ d3 r ⃗ exp [i(K⃗ − K⃗ ) ⋅ r]⃗ = ∫ d3 r ⃗ exp [i(K⃗ − K⃗ ) ⋅ (r ⃗ + d)] (1.46)
V a Va
This expression must be satisfied by any d,⃗ however, it is also correct for K⃗ ≠ K⃗ pro-
vided that the integral vanishes:
Thus, we have proved that the identity (1.45) holds true for K⃗ ≠ K⃗ . Moreover, it remains
valid for K⃗ = K⃗ as well, because (1/V a ) ∫V d3 r ⃗ = 1.
a
Using (1.45), it is easy to verify the validity of the representation (1.44) for the co-
efficients f K⃗ in the expansion (1.43). To this end, we should multiply the equality (1.43)
by exp(−i K⃗ ⋅ r)⃗ and integrate the result obtained over r ⃗ within the unit cell of V a .
ℎ2
− ⃗ = εΨ ,
∆Ψ + U(r)Ψ (1.48)
2m
24 | 1 Electrons and Holes in Metals and Semiconductors
the potential energy has the periodicity of the Bravais lattice: U(r ⃗ + R)⃗ = U(r),
⃗ so we
represent the function U(r)⃗ in the form of a Fourier series:
where K⃗ is the reciprocal lattice vector. The Fourier coefficients U K⃗ are related to U(r)⃗
by:
1
U K⃗ = ∫ d3 r ⃗ exp (−iK⃗ ⋅ r)⃗ U(r)⃗ . (1.50)
Va
Va
In formula (1.50), the integration is over the elementary unit cell V α of the crystal’s
Bravais lattice.
2. We can always replace the potential energy by an additive constant value. Then
we choose the value so that the mean value of the potential U0 , taken for one elemen-
tary unit cell, should vanish:
1
U K=0
⃗ = ⃗ r)⃗ = 0 .
∫ d3 rU( (1.51)
Va
Va
holds true.
4. When a crystal has a central symmetry, a suitable choice of the coordinate sys-
tem yields: U(r)⃗ = U(−r).
⃗ Then (1.50) and (1.52) imply the chain of equalities:
U K⃗ = U−K⃗ = U ∗⃗ . (1.53)
K
Thus, for crystals with an inversion center, U K⃗ is the real quantity. Note that the one-
dimensional lattice always has a center of inversion (see Figure 1.15).
Imagine that the sample is in the shape of a parallelepiped (Figure 1.14), built on
the vectors L⃗ i = N i a⃗ i . Here i = 1, 2, 3. It is worth pointing out that there is no summation
N3 a 3
L=Na
N2 a 2
N1 a 1 a
Fig. 1.14: A parallelepiped shaped Fig. 1.15: The chain of N atoms is a one-
crystal. dimensional crystal.
1.5 Electrons in a Weak Periodic Field | 25
over the twice repeated index i. N i are positive integers that define the lengths of the
sides of the parallelepiped.
For U(r)⃗ = 0 and the periodic Born–Karman conditions:
Ψ ( r ⃗ + L⃗ i ) = Ψ(r)⃗ , (1.54)
1 ℎ2 k⃗ 2
Ψ k⃗ = exp (ik⃗ ⋅ r)⃗ , ε n (k)⃗ = .
√V 2m
Here V = L⃗ 1 ⋅ [L⃗ 2 × L⃗ 3 ] is the volume of the crystal; the allowed values of the wave
vector have the form:
3
mi ⃗
k⃗ = ∑ bi , (1.55)
N
i=1 i
where m i are arbitrary integers. For the sample’s large sizes (for large N i ), the wave
vectors (1.55) completely fill the entire k-space.
Our goal is to use perturbation theory to calculate changes of wave functions and
energy levels of an electron in a weak periodic field U(r)⃗ ≠ 0. Calculations and final
results for three-dimensional and one-dimensional crystals are close. Certainly, the
one-dimensional case is atypical in many respects; it differs from the two-dimensional
and three-dimensional ones. For example, the one-dimensional case does not imply
the introduction of the Fermi surface – there are no overlapping bands. However, this
problem makes complex geometric concepts more pictorial. Therefore, we give cal-
culations and figures for the one-dimensional crystal. And then we come back to the
three-dimensional problem and discuss a real situation.
A one-dimensional crystal is a chain of N atoms of length L = Na, where a is the
distance between neighboring atoms (Figure 1.15).
The normalized wave function Ψ k and levels of the energy ε0 (k) of a free electron
in the one-dimensional crystal can be written as:
1
Ψ k (x) = exp (ikx) ,
√L
(1.56)
ℎ2 k 2 s 2π
ε0 (k) = , k ≡ ks = .
2m N a
The function (1.56) satisfies the Born–Karman boundary condition Ψ(x + L) = Ψ(x)
if s is an arbitrary integer.
The energy of a free particle as a function of the wave number has the form of
a parabola (Figure 1.16). The dots in the parabola are the allowed values of the elec-
tron energy. For large N, the distances ∆k between adjacent points along the axis k
decrease; ∆k = (2π)/(Na) → 0 as N → ∞. Therefore, when it comes to the electron
energy, we will often draw continuous curves, ignoring the fact that the curves are sets
of closely spaced discrete points.
26 | 1 Electrons and Holes in Metals and Semiconductors
ε(k)
The periodicity condition for the one-dimensional crystal has a simple form:
U(x + na) = U(x) ,
where n is an arbitrary integer. For the wave vectors of the reciprocal lattice: K ≡ K s =
2πs/a, s is the integer.
Next, we expand the potential U(x) in a Fourier series in a fashion that facilitates
the further passage from the one-dimensional to three-dimensional case:
2π
U(x) = ∑ U K exp (iKx) = ∑ U K n exp (i nx) ,
K n a
a
1
U K = ∫ dx exp (−iKx) U(x) .
a
0
To make allowances for the energy ε0 (k) of the free electron due to the periodic
potential U(x), we need to use quantum mechanical perturbation theory for the sta-
tionary Schrödinger equation:
ℎ2 d2 Ψ
− + U(x)Ψ = εΨ .
2m dx2
Considering the weak potential U(x) as a perturbation, in accordance with the pertur-
bation theory we have:
̂ 2
̂ ⟨Ψ k U Ψ k ⟩
ε(k) = ε0 (k) + ⟨Ψ k U Ψ k ⟩ + ∑ + ⋅⋅⋅ , (1.57)
k ,k =k
̸
ε0 (k) − ε0 (k )
where |Ψ k ⟩ are the vectors of a free electron state.
Let us estimate the matrix elements ⟨Ψ k |U|Ψ ̂ k ⟩ for the one-dimensional crystal
(k ≡ k s , k ≡ k s ):
L=Na
̂
⟨Ψ k U Ψ k ⟩ = ∫ dxΨ∗k U(x)Ψ k =
s s
0
L=Na
1 2πix 2πin
= ∫ dx exp [ (s − s)] ∑ U K n exp ( x) =
L Na n a
0
L=Na
1 2πix
= ∑ ( ∫ dx exp [ (s − s + nN)]) U K n . (1.58)
L n Na
0
1.5 Electrons in a Weak Periodic Field | 27
A similar outcome holds in the three-dimensional case. The passage to the three-
dimensional crystal corresponds to the formal replacement: k → k,⃗ k → k⃗ , K → K⃗
in (1.59).
We have chosen the point of reference for the potential energy of the crystal so that
the matrix element ⟨Ψ k |U|Ψ ̂ k ⟩ = U0 vanishes (1.51). Therefore, the first nonzero cor-
rection to the energy of the free electron due to the periodicity of the potential energy
appears as:
̂ 2
⟨Ψ k U
∑ Ψ k ⟩ = ∑ |U K |2
.
ε0 (k) − ε0 (k ) ε0 (k) − ε0 (k − K)
k ,k =k
̸ K=0
̸
|U K |2
ε(k) = ε0 (k) + ∑ . (1.60)
K=0
̸
ε0 (k) − ε0 (k − K)
The perturbation theory has a right to exist when each subsequent term of the expan-
sion is less than the previous one. In this given case, this requirement is not met in
the region of the point k = K/2 because ε0 (K/2) = ε0 (−K/2) and the denominator in
the second summand in (1.60) becomes small for k ≈ K/2. For such values of k, the
perturbation theory is inapplicable and in need of modification.
For the three-dimensional case, calculations and a result are analogous to the one-
dimensional case:
2
U K⃗
ε(k)⃗ = ε0 (k)⃗ + ∑ .
⃗ ε0 (k)⃗ − ε0 (k⃗ − K)⃗
K=0
̸
The three-dimensional case differs from the one-dimensional one in that k⃗ is a vector
rather than a number. Hence, the condition (1.61) gives a constraint not only to the
magnitude but also to the direction of the wave vector.
To analyze the condition (1.61), we need to write it explicitly:
ℎ2 k⃗ 2 ℎ2 ⃗
= (k − K)⃗ 2 .
2m 2m
28 | 1 Electrons and Holes in Metals and Semiconductors
A plane in k-space
O
k k k K
K
k K
K
K2
a) |k|⃗ = |k⃗ − K|⃗ , b) k⃗ ⋅ K⃗ = . (1.62)
2
It is known from analytical geometry that, for a fixed K,⃗ a plane perpendicular to the
vector K⃗ in k-space corresponds to equation (1.62) b. In the reciprocal space, the wave
vector k⃗ of an electron ties the starting point O to any other point of this plane. From
equation (1.62) a, it follows that the plane is not only orthogonal to the vector K⃗ but it
also accurately divides it in half. It is easy to see that the last assertion is true by using
the hypotenuse leg theorem (theorem of the equality of right-angled triangles with the
same cathetus and hypotenuse) (Figure 1.17).
Recall that K⃗ is a vector of the reciprocal crystal lattice. Suppose K⃗ is the shortest
in a discrete set of parallel vectors of the reciprocal crystal lattice. This would make
a plane passing through its middle the first Brillouin zone boundary. As a result, the
perturbation theory is inapplicable, for example, near the boundaries of the Brillouin
zone.
Inside the first Brillouin zone ε0 (k)⃗ ≠ ε0 (k⃗ − K/2),
⃗ the perturbation theory consid-
ered here is valid. The energy levels of the electron in the weak potential differ little
from the energy levels of a free electron. This is one of the reasons why the first Bril-
louin zone is the most convenient primitive unit cell in reciprocal space to determine
the electron wave vector.
When lengths of the vectors K⃗ in reciprocal space are arbitrary, there are, except
for the of the first Brillouin zone boundaries, other planes of the type:
k⃗ ⋅ K⃗ = K 2 /2 . (1.63)
The theory of perturbations can also be utilized for them. Let us elucidate the physical
meaning of the condition (1.63).
It would be appropriate at this point to recall that the electron wave vector is re-
lated to the de Broglie wavelength, correlated to the electron: |k|⃗ = 2π/λ. Let K⃗ min be
Exploring the Variety of Random
Documents with Different Content
“Cart-loads,” said he, absently, wondering what had caused her
augmented color, and watching her as he would always now watch
her whenever there was the slightest deviation from her normal
manner.
“And I suppose,” she said, “she spends a great deal on them?”
“I suppose so,” he answered, “judging by the number that I’ve
seen wither in their prime and disappear, and new ones take their
places the next day.”
Viola pressed the lock in and shot it out.
“Are any of them dead just now?” she asked, in rather a small
voice.
“Dozens, probably. It seems to me some of them are always
dead, only they’re considerate enough not to all die at the same
time.”
There was a moment’s pause. Gault’s gaze was diverted from her
face to the high, old-fashioned room, with its marble mantel carved
in fruits and flowers and its bare sideboard. Then Viola said:
“Your sister-in-law always gets her plants from the large florists,
doesn’t she? Some one on Kearney or Sutter Street?”
“I dare say she does; but I’m sure I don’t know. I can’t control
my curiosity any further—what were you going to do with those
round bits of paper you were cutting when I came in?”
She looked at him quickly, a look of sharp, dubious inquiry; then,
as she met the amused curiosity of his glance, she gave a little laugh
and said:
“I was going to make jam.”
“But you don’t make jam out of paper?”
“No; those are for the tops of the glasses. I soak them in brandy
and put them on, and they preserve it.”
He looked at the papers, then back at her. As their eyes met the
delight each felt in the other’s presence found expression in a
simultaneous burst of laughter. For a moment they stood facing each
other, laughing in foolish but happy lightness of heart.
“Now, you know,” he said, “I’m a credulous person, but isn’t that
going too far? Why, if you used all those things you’d have jam
enough to feed the American army.”
Her laughter died, and looking slightly confused, she put out her
hand, seized the other door, and drew them together with a bang.
“There!” she said, dropping the catch; “you can’t see any more.
You’re too curious, in the first place, and you don’t believe me,
which is worse.”
“I’ve found out the skeleton in the closet,” he said, as they
walked back into the front room. “It’s the colonel’s passion for jam.
I’ve heard of a passion for pie running in families, but jam’s
something new.”
The bare austerity of this bleak apartment seemed to cast a
sudden chill over their high spirits. Gault, sitting in the colonel’s
chair, reverted in thought to the object of his visit, and wondered
how he could turn the conversation in the direction he had intended.
His preoccupation, and the sense of shame he felt at the mean part
he contemplated playing, made him respond to her conversational
attempts with dry shortness. She grew constrained and
embarrassed, and finally, in a desperate attempt to arrest a total
silence, said:
“Don’t you like my new cushion? You’ve never noticed it!”
The visitor’s slow glance moved in the direction indicated, and
rested on a cretonne cushion in one of the wicker chairs.
“It’s a perfect beauty,” he said, with as much enthusiasm as he
thought the occasion required.
“I’m glad you think it’s pretty,” she answered, evidently much
pleased. “I ought not to have bought it, I suppose, but I do love
pretty things.”
“Why oughtn’t you to have bought it? What is the matter with
it?”
“Nothing; I mean it was an extravagance. I sometimes think how
perfectly delightful it would be to be able to go into stores and buy
furniture and ornaments and curtains just whenever you wanted.”
This remark dispelled Gault’s preoccupation. He remained in the
same position and continued staring at the cushion, but his glance
had changed from its absent absorption to a fixed and listening
intentness.
Viola saw that she had interested him, and continued with happy
volubility:
“Sometimes, when I have nothing to do and am here alone, I
think how I would furnish this room if I could buy anything I saw,
and could just say to some outside person, the way princesses do, ‘I
have bought so much; please pay the bill.’ I’ve done it in white and
gold, and in crimson with black wood, teak or ebony, very plain and
heavy; and also in striped cretonnes with bunches of flowers, and
little chairs and sofas with spindle legs. There’s a great deal of
satisfaction in it. It’s almost as good as having it really happen.”
“It sounds very amusing,” said Gault, as she paused; “but then,
castles in the air,” he added, turning to look at her, “are never quite
the same as the real thing.”
“If you can’t get the real thing, you take the castles in the air,”
she answered, smiling.
“Tell me some more of yours.”
“Oh, they’re just silly dreams, and mercenary ones, too. My
castles are all built on a foundation of money. It’s a dreadful thing to
have to acknowledge, but I’m afraid I am mercenary. And it’s such a
horrid fault to have.”
“But isn’t it rather a useful one?” he could not forbear asking.
“Not so far. Once I had my palm read by a palmist, and he told
me I was going to be very prosperous—to have great riches. That’s
one of my best castles in the air. I’m all the time wondering about it,
and where my great riches are coming from.”
She spread both hands, palms up, on the table, and studied
them as if trying to elicit further secrets from their delicately lined
surfaces.
“Great riches!” she repeated. “Where could a person suddenly
find great riches? The mining booms are over, and in California
people don’t strike oil-wells in their gardens. I’m afraid it will have to
be either begging, borrowing, or stealing. I wonder which I would
succeed best in.”
With the last words she raised her bent head, and her eyes,
diminished in size by her laughter, rested full on his. Their glance
was clear, candid, and innocently mirthful as that of a merry child.
As he stared at her, almost vacantly, the notes of a clock, striking
somewhere in the back of the house, fell with crystalline distinctness
upon the silence.
“One—two—three—four—five,” she counted absently, with each
number touching the table with a finger-tip.
Gault rose to his feet, remarking with unfeigned surprise on the
lateness of the hour. She looked suddenly confused and annoyed at
the realization of her unintentional rudeness, and asked him if he
would not remain till her father’s return. But he pleaded an
engagement he had made to attend the tea given that afternoon by
Mrs. Jerry McCormick, and, with a hand pressure and the
conventional words of farewell, brought his visit to a close.
Outside, he turned to the right and walked slowly forward toward
where the rumble of traffic indicated one of the large and populous
thoroughfares of the district. Before him, at the end of the street’s
long vista, the sunset glowed pink, barred by a delicate scoring of
telegraph-wires. Even as he looked it deepened and burned higher
and higher up the sky, while at the far end of the vista it
concentrated into a core of brightness, as though a conflagration
were in progress there.
What was he to think? He felt his mind confused and full of
warring images. He had been almost afraid of what she might say—
she who was to him the ideal of all that was gentlest and truest and
most maidenly. And yet what had she said to disturb or annoy him?
It was only the foolish prattle of a girl who is happy and in high
spirits. And even as he made these assurances to himself, sentences
from the past interview surged up to the surface of his mind: “I’m
afraid I’m mercenary, and it’s such a horrid fault to have.” “Where
are my riches coming from? It will have to be either begging,
borrowing, or stealing.”
Her mother had been an actress—one of the stars of San
Francisco’s hectic youth. Dissimulation might be instinctive with a
woman of Viola Reed’s heredity. It was the whole art of acting; it
was in her blood. He thought of all he had ever heard of her mother,
of her few years of fame and glory, so splendidly ended by her
marriage to the bonanza millionaire. It had been a wonderful,
glittering life, quenched in an early death. He had never heard
anything against her character, but she had been an actress, the
essence of whose art is the capacity to both conceal and assume
emotion. And her daughter, in personal appearance at least,
resembled her. He had heard that from the colonel himself.
A feeling of weariness and disillusion took possession of him, and
in the sickness of heart that it brought he thought suddenly of
Letitia. She was the one woman he knew that he could always rely
on to be true and steadfast and genuine. Why had he not loved her
—a woman a man could trust forever, and handsome enough to be
the wife of a king? There would be no doubts nor difficulties in a life
with her; it would be all kindness and cheer and sympathy. And even
as he thus reflected, he knew that love for Letitia was as far from
him as was indifference to the woman whom he mistrusted.
At the very hour that Gault was walking moodily across town
from South Park, Letitia, the object of his thoughts, was rolling along
the asphalted streets of the Western Addition in Mrs. Mortimer
Gault’s coupé. Her sister was with her, and both ladies were
dressed with a rustling splendor which betokened festal doings. For
they, too, were en route to the McCormick tea. This was, in fact, a
large reception given by Mrs. McCormick to little Prince Dombroski, a
gentleman who had come from Russia to wed a Californian heiress,
and was receiving a helping hand from the McCormicks, who on this
particular afternoon had gathered together all maiden and widowed
San Franciscan wealth for his inspection.
Letitia had dressed herself for the occasion with great care.
When she had appeared at the front door and descended the stairs
to the carriage, she had presented so dazzling a picture that even
the coachman, a well-trained functionary imported from the East,
could hardly forbear staring at her. She was regally clothed in a
costume of bluish purple, with much yellow lace, fur, cream-colored
satin, and glints of gold braiding about the front. There was a purple
jewel at her throat, and a bunch of pale, crape-like orchids, that
toned with the hue of her dress, was fastened on her breast. Clad
thus in the proudest production of a great French modiste, Letitia
was really too handsome to be quite in good taste. But she was used
to sumptuous apparel, and carried it with the air of an actress who
knows how to take the stage.
Maud Gault was somewhat less punctual to-day than her sister.
Letitia sat in the carriage waiting for her, and finally, by the brushing
of silken skirts and an advancing perfume of wood-violet, was
apprised of her sister’s approach. The elder woman gave the address
to the coachman and then sprang in.
Hardly had the door closed when she looked at Letitia with a
kindling eye, and said:
“Oh, Tishy, I know the funniest thing!”
Letitia knew that her sister had something of note to impart. Mrs.
Gault’s dark cheek was flushed a fine brick-red, her eye was alight.
She was pulling on her gloves as she spoke.
“Do you remember that night, only a few weeks ago, when you
asked John about Colonel Reed’s daughter?”
“Yes.”
“And do you remember that he said he’d never seen her?”
“No, he didn’t say that,” corrected Letitia; “he said he’d heard of
her.”
“And what else?” asked the other, stopping in her glove-pulling to
fix Letitia with a keen eye.
“I don’t think he said anything else. I don’t remember anything.”
“But he certainly led us to believe that he didn’t know her. Didn’t
he, now?”
Letitia paled slightly. Her eyes, looking frankly troubled, were
fastened on her sister.
“Yes—I think so. Why?”
“Well, my dear,” said Mrs. Gault, bridling with the consciousness
of her important announcement, “he knows her well. He goes there
all the time. He’s having a regular affair with her. Did you ever know
anything to beat men?”
“How do you know?” said Letitia, looking down and picking at the
gold arabesques on her dress.
“Mortimer told me last night. He made me swear I wouldn’t tell a
living soul. You must remember that, by the way, or I’ll get into
trouble. Mortimer saw Colonel Reed in the office the other day, and
that red-haired clerk, the one John took in because his mother was
crazy or consumptive or something, told Mortimer Colonel Reed
came there often, and that John went out to see him at his home
somewhere near South Park. Doesn’t that beat the band? John going
calling in South Park on Colonel Reed’s daughter, and then
pretending to us that he doesn’t know her! If John knew the man
had said anything about it, he’d kick him down all the stairs in the
building, if they reached from here to the ferry.”
Letitia was silent. She thought of the conversation on Sunday,
and the woman who had been the heroine of the novel. All the
sunshine seemed to go out of the afternoon, and the innocent joy
she had taken in putting on her beautiful clothes suddenly shriveled
up and vanished.
“He might go out there and see Colonel Reed’s daughter and not
tell us about it,” she said, “and yet not—not be exactly in love with
her.”
“Dear me, Letitia,” said her sister, pettishly, “what a dunce you
are! Do you suppose John’s going to drag himself over to South Park
to see Colonel Reed’s daughter because he’s taken a philanthropic
interest in her father? One would think you’d been raised in Oshkosh
or Milpitas, to hear the things you sometimes say. But that’s not all.
This morning I was in the Woman’s Exchange, and who should be
there but old Biddy McCormick herself. I can’t endure her, you know,
especially since she’s got this little prince-creature up her sleeve; but
I’m always polite to her because of Tod and you—and things
generally. You never can tell what may happen. And I heard her say,
‘Not that jam; I always buy the same kind—Miss Viola Reed’s.’ So I
up and said, as innocent as Mary’s little lamb, ‘Do tell me, Mrs.
McCormick, what jam that is you’re buying. Everything you have is
always so delicious.’ And she said, ‘It’s some that’s made by a
woman named Reed, who lives across town somewhere.’ Then,
when she’d gone, I corralled the girl, and she told me it was made
by a Miss Viola Reed, who lives—”
Mrs. Gault opened her jeweled card-case and produced a slip of
paper with an address written on it. She handed this to Letitia, and
said with an air of triumph:
“That’s where she lives. Now you’ll have to admit, Miss Letitia
Mason, that there are no flies on your little sister!”
Letitia looked at the address and gave it back.
“No,” said her sister; “you keep it. That’s my little scheme. You’re
to go there now—this afternoon—and order jam. Do you see?”
“But I don’t want any jam, and you never eat it.”
“Good gracious, Tishy, how awfully stupid you are to-day! What a
fortunate thing it is that you and Mortimer have got me to take care
of you! Of course you don’t want jam. I never heard of any civilized
being who did. But I suppose you’ll admit that you want to see this
girl?”
“I don’t think I do,” said Letitia. “I don’t see why I should.”
“Well, I do,” said Mrs. Gault, with asperity. “Don’t you take an
interest in John? Don’t you want to see if he’s fallen into the clutches
of an adventuress?”
“She doesn’t sound at all like an adventuress, Maud. I never
heard of an adventuress making jam for her living.”
“Jam for her living! Bosh! Can’t you imagine how she tells that to
John, and shows him the glasses in the corner cupboard, and lets
him find her stirring things in a big pot on the kitchen stove? Oh,
she’s no fool, my dear! Will you go and see her?”
“I’d rather not.”
“Very well, then; if you care so little for John, you needn’t go. I’ll
do it myself, and I can tell you, I’ll size her up.”
Letitia looked uneasy. She knew nothing of Miss Reed except that
she was poor and pretty. But she did not like the thought of
subjecting even an unknown female to Mrs. Gault’s mercies, when
her interest was so evidently hostile and her curiosity so poignant.
“If you think somebody must go, then I will,” she said pacifically.
“I don’t see the use of it, but I can go better than you.”
“All right,” said Mrs. Gault, immediately placated. “You’d better go
now. It’s always best to do a thing when you have the opportunity.”
“No,” said Letitia; “I don’t think I’ll do that.”
“Why not? Is it possible you’re so crazy to see that miserable
little prince that I could put in my hat-box?”
“I don’t care about him,” answered the girl, with unmoved
placidity. “I don’t like to go—to go this way.” She made an
explanatory gesture toward her dress.
Mrs. Gault looked at her uncomprehendingly.
“Why? What’s wrong about your clothes?”
It was painful, but Letitia had to explain:
“If she’s so poor as all that—and everybody says so—I don’t think
it’s—it’s—quite nice, some way or other, for me to go in this dress.”
Her voice took on a sudden tone of decision. “I won’t do it, anyway.”
Her sister knew the tone, and knew that there was no use in
combating the mood it indicated.
“You have the queerest notions,” she said, with a resigned sigh;
“but do as you like. It’s all the same, if you do go to-morrow. Only
you must promise that you won’t back out.”
Letitia promised.
On the afternoon of the next day she stood before her glass and
critically eyed her reflection. She had put on a plain tailor-made suit,
which fitted her heavily molded figure with unwrinkled smoothness.
A brown turban crowned her reddish hair, and the exquisite pallor of
her skin was obscured by a thin veil. Letitia did not approve of
herself in this modest garb. She accepted the dictum that “beauty
should go beautifully.” But for the mission upon which she was
bound she had selected her attire with an eye to its fitness and
propriety.
It was a gray afternoon, with a breath of fog in the air. Already
the city was beginning to show signs of the summer exodus, and
Letitia was glad that in her journey across town she met no
acquaintances and attracted no more attention than that frankly
candid stare which is male California’s passing tribute to beauty.
Though she had been born in South Park, she knew nothing of
this side of the city, and found herself as much a stranger as its
inhabitants would have been had they been transported to the
aristocratic heart of the Western Addition. Finally, however, after
some questioning of small boys and much retracing of steps, she
found the house, and walked up the path with the black-and-white
flagging.
Letitia was one to whom the word “shyness” has no meaning.
She possessed her full share of the Westerner’s placid self-approval,
and with it that careless curiosity which makes an incursion into new
surroundings interesting. Yet, as she stood waiting for the door to
open, she experienced a sensation of nervousness quite new to her.
Her heart had ached more in the last twenty-four hours than it had
since her mother’s death, years before. If Viola Reed was an
adventuress or if she was a saint, the situation was equally painful to
this splendid-looking creature, who, for all her regal air and stately
immobility of demeanor, was only a woman of a simple, almost
primitive type.
The door was opened by Viola, in her blue gingham dress and
her apron. At the sight of her visitor she looked startled almost into
speechlessness. Letitia announced the fact that she had come on
business, and an invitation to enter brought her sweeping into the
little hall and the drawing-room beyond.
Here the two girls looked at each other for one of those swift
exploring moments in which women seem to take in every detail of
dress, every peculiarity of feature and revealing change of
expression, that a rival has to show. Letitia, with all her apparent
heaviness, had keen perceptions. With a sinking at her heart she
saw the beauty of the gray eyes fastened shyly upon her, and
realized what must be the power of the delicate charm, so far
removed in its soft, dependent femininity from her own. She saw
that this girl had a distinguishing refinement she could never boast,
and that it was strong enough to triumph over such poverty-stricken
surroundings as, in all her experience, she had never before
encountered. Her quick eye took in the gaunt emptiness of the room
as John Gault’s could not have done in a week’s arduous
examination. She saw the split and ragged shades in the windows,
the ribs of twine in the old carpet, the rents in the colonel’s chair.
Viola, for her part, saw one of the handsomest and most
imposing young women she had ever gazed upon. The very way
Letitia rustled when she moved, and exhaled a faint perfume with
every movement, seemed to breathe an atmosphere of fashion and
elegance. She had never seen her before, and had no idea who she
was. Letitia soon put an end to this condition of ignorance.
“My name is Mason,” she said judicially—“Letitia Mason. I am the
sister of Mrs. Mortimer Gault.”
At this announcement an instantaneous change took place in
Viola. For a second she looked alarmed, then her face stiffened into
lines of pride and anger. The eyes that had been so full of a naïve
admiration were charged, as by magic, with a look of cold
antagonism. Letitia felt her own breath quicken as she realized how
much the name of Gault must mean to this girl.
Viola attempted no answer to the introduction, and Miss Mason
hastily went on:
“My sister heard that you made jam—very good jam. We don’t
like what we get in the stores, so we thought we would try yours.”
Viola had now found her voice,—a very low and cold one,—and
answered:
“You can get it at the Woman’s Exchange. I sell it there all the
time.”
“Yes, I know that,” said Letitia; “but we thought it would be
better to buy it straight from you; that—perhaps—it—perhaps it
would save time and trouble.”
“I don’t see how it could do that. This part of town is a long
distance out of everybody’s way.”
“Yes, of course it is,” the other agreed eagerly; then, with a
sudden happy inspiration, “but I thought you might have a larger
variety here—that you might have a good many different kinds on
hand. I don’t want all the same sort.”
Viola rose and went to the door that led to the dining-room. Her
resentment was not more obvious than her embarrassment. There
was something tremulous in the expression of her face that gave
Letitia a wretched feeling that only pride enabled her to keep back
her tears.
“I have just the same here that I have at the Exchange,” she
said, opening the doors.
The visitor followed her. In the gray of the afternoon the long
room, with its tiers of plants and its bare sideboard and mantelpiece,
looked even colder and drearier than the drawing-room. Viola
opened a cupboard and indicated the lines of glass jars standing on
the shelves. She tried to be businesslike, and told their contents and
prices, but her voice betrayed her. Letitia, listening to her and staring
at the Chinese cracker-jar that was the sole adornment of the
sideboard, suddenly felt sick with disgust at herself for intruding, at
her sister, at John Gault.
As Viola’s voice went on,—“These are apricots; they’re fifty cents.
Those on that shelf are strawberry and raspberry; they are only
thirty,”—Letitia’s shame and indignation worked up to a climax and a
resultant resolution.
She took up one of the glasses and, looking at the legend written
in neat script on the paper top, said:
“I think I ought to tell you how I happened to come here. It’s
really a secret and you mustn’t tell. What I said at first was not quite
the case. No one at our house knows anything about this but me.
I’m going to buy these preserves for my brother-in-law and tell him I
made them. I’m going to fool him. Do you understand? It’s just a
little joke.”
Letitia delivered herself of this amazing effort at invention with
admirable composure, for it was the first elaborate and important
falsehood she had ever told in her life. Viola, turning from her
contemplation of the shelves, looked at her, relieved but not quite
comprehending.
“So I hunted you up myself at the Exchange,” continued Letitia,
plunging deeper into the slough of deception, but knowing now that
she had gone too far to compromise with truth, “and came here
myself this way so as to keep it all dark.”
Viola’s face had cleared with each word. As the other ended, her
lips parted in the smile that John Gault found at once so irresistible
and so enigmatic. Letitia found nothing enigmatic in it. She only
thought, with a piercing dart of pain, “She is still prettier when she
smiles.”
“It’s very amusing,” said Viola; “but why do you want to fool
him?”
Letitia was even ready for this, so expert does the first lie make
us in perpetrating the second.
“He says I am useless and can’t do anything. I am going to show
him that I can make jam.”
Viola was rather shocked, but relief and amusement combined to
make her light-hearted, and this time she laughed.
“But the writing,” she said. “Won’t he see by that that it’s not
yours? There’s writing on every glass.”
“Oh, that will be all right. I’ll have the Chinaman put it out in a
dish. But you’ll promise not to give me away?”
“Oh, I never will,” said Viola. “In fact,” she continued naïvely,
“I’d rather have it that way myself. You see, many people—all
people, that is—don’t know that I do this.”
She stopped and looked tentatively at Letitia, as if curious to see
how she was taking these revelations.
“Do what?” asked Letitia, not understanding.
“Make the jam. Not that I mind much. But it’s a little sort of fancy
of my father’s. Sometimes older people have those ideas, and it’s
best to humor them, I think; don’t you?”
“Oh, much the best,” assented the other, turning aside and
looking at the plants. “It’s best to humor everybody; it’s so much
easier to get on. What beautiful ferns!”
“Yes; I am quite proud of them. But this is a splendid window for
ferns.”
“Did you raise these yourself? I never saw such plants out of a
greenhouse.”
Viola was now eagerly interested.
“Yes, I grew them all—some of them from a few roots like black
threads. I sell these, too. There is a man at one of the Kearney
Street florists’ who used to live near here and knew us, and he buys
them from me. At Christmas I do quite well.”
Letitia examined the ferns.
“I wonder if you would let me buy one or two of them,” she said.
“We can’t get such plants at our florist’s, and I am fonder of them
than of any other kind of fern.”
Viola agreed with a blush of pleasure, and after some
consultation four ferns were selected. The visitor was amazed at
their cheapness, but concealed her astonishment. Then she bought
three dozen jars of the jam. She did not pay for them, but said that
on the following day she would send the money by a messenger,
who would also bring away the purchases.
Standing in the doorway, about to leave, she said:
“I’m glad to have seen you. It’s so interesting for a person like
me, who can’t do anything, to meet some one who is clever and of
use in the world. Good-by!” She held out her hand, and Viola,
surprised, put hers into it. “Don’t forget to keep our secret. It makes
a person feel like a conspirator, doesn’t it? I think, too, Colonel
Reed’s quite right to want to be reticent about business matters. So
you and I’ll keep dark about this little transaction of ours.”
This was the most diplomatic sentence Letitia had ever given
vent to in her life.
She walked slowly away from the house, her eyes downcast in
thought. The superb health she had inherited from an untainted
peasant ancestry made her imagination dull, and lightened such
sufferings as she had encountered in her easy, care-free life. Even
now she experienced none of those fierce pangs that jealousy and
disappointed love provoke in the women of a more sophisticated
stock. She was made on that large, calm plan on which an all-wise
nature creates the maternal woman—she whose destiny it is to bear
strong children to a stalwart sire. But this afternoon, for the first
time in her life, she knew what it was to feel her heart lying heavy in
her breast like a thing of stone.
It was late when she reached home. Mrs. Gault was to give a
dinner that evening, and as Letitia passed through the hall she
caught a glimpse of her sister, in a loose creation of pink silk and
lace, which swelled out behind her like a sail, hurrying round the
bedecked dining-table, followed by two meek and attentive
Chinamen. Knowing the indignation of Maud should she be late, she
ran to her room and made her toilet with the utmost speed.
She was just completing this important rite, and, seated at her
dressing-table under a blaze of electric light, was selecting an aigret
for her hair, when the door opened and Mrs. Gault entered.
She had discarded her ebullient draperies of pink silk, and was
sheathed tightly in her favorite yellow, from which the olive skin of
her bared neck emerged in polished smoothness. As she came
forward she had one hand full of diamond brooches, which she
pinned with apparent carelessness round the edge of the low bodice.
“Well, Tishy,” she said, sitting down by the dressing-table, “what
happened?”
Letitia looked at the array of silver that covered the table. Some
jewels lay scattered among it, and the aigrets from which she had
been about to choose the one she should wear. She selected a black
one, and turned it round, looking at it.
“Nothing happened,” she answered. “I saw her, and bought the
jam and some plants. She raises plants, too.”
“Is she really so pretty?”
“Yes, very—I think some people might say beautiful.”
Mrs. Gault’s face fell.
“She didn’t say anything about John, I suppose?”
“Of course not.”
“Did it seem to you that there was anything adventuressy or bad
about her?”
Letitia looked at her sister—a sidelong look, which made Mrs.
Gault feel rather uncomfortable.
“I never saw any one in my life that looked to me less so,” she
answered.
“Dear me!” ejaculated Mrs. Gault, in a dismayed tone. “You don’t
say so! Tishy, for goodness’ sake, look where you’re putting that
aigret! You look like Pocahontas, and Tod McCormick’s coming to
dinner.”
Letitia arranged the aigret at a more satisfactory angle, her large
white arms, shining like marble through the transparent tissue of her
sleeves, shielding her face.
“Then,” said Mrs. Gault, returning to the more important subject,
“there really may be a chance of his marrying her.”
“I should think a very good one,” answered Letitia, in a low voice.
“Good heavens!” breathed her sister, in the undertone of utter
horror, “how awful men are! What makes you think he may intend
marrying her?”
“Because,” said Letitia, dropping her arms and turning on her
sister with her mouth trembling and her breast agitated with sudden
emotion, “no man who was any sort of a man could mean anything
else.”
Maud Gault was amazed by the girl’s unexpected emotion. She
pushed back her chair, and staring at Letitia, said vaguely:
“Why? I don’t understand.”
“Even if he didn’t care, even if he didn’t love her, he’d marry her.
Oh, Maud, she’s so helpless and so poor!”
And Letitia burst into a sudden storm of tears.
For a moment her sister sat still, looking at her in blank
amazement. Then she felt a pang of feminine sympathy. So Letitia
did care for him. Poor Tishy!
“There, don’t cry!” she said, patting her shoulder. “You never can
tell about these things. John may not care a button for this girl, or
have the least intention of marrying her. You’re always seeing the
dark side of things.”
But her form of consolation was not well chosen. Letitia threw off
the hand and raised her disfigured face.
“John may be selfish and mean and all that, and I’ve no doubt he
is; but he’s not mean enough, he’s not contemptible enough, to do
what you think he’s doing. I’ll not believe that of him. I’d despise
him if I thought so; I’d hate him!”
Her tears burst forth afresh, and she hid her face in her hands.
Mrs. Gault was nonplussed. She looked at her sister’s shaken
shoulders and bowed head with an uncomprehending but pitying
eye. Then, as Letitia’s sobs diminished, she said gloomily:
“How much jam did you buy?”
“Three dozen glasses,” came the muffled answer.
“Good gracious!”—raising her eyes toward the ceiling in an
access of horror. “What did you get so much for? Two or three would
have done. We’ll not get through that by Christmas.” There was no
answer made to this, and after a moment or two of silence Mrs.
Gault recommenced, in a brisk and unemotional tone:
“I don’t understand you at all, Tishy; but I do know that if you
don’t stop crying you’ll look a perfect fright at dinner, and everybody
will be wondering what’s the matter with you.”
This appeal to her pride had a good effect upon Letitia. She
struggled with her tears and finally subdued them. But her flushed
and swollen countenance needed much attention, and when Mrs.
Gault left the room she carried with her a picture of her sister sitting
before the mirror solicitously dabbing at her eyelids with a powder-
puff.
When she appeared all traces of her previous distress seemed
successfully obliterated. It remained for the eye of love to penetrate
the restorative processes with which she had doctored her telltale
countenance.
Near the end of dinner Tod McCormick, who sat beside her,
leaned toward her and said, in the low tone of long-established
friendship:
“What’s the matter, Tishy? You look sort of bunged up.”
Letitia said nothing was the matter—why?
The small, red-rimmed eyes of Tod passed over her face,
lingering with the solicitude of affection upon the delicately pink
eyelids and nostrils.
“You look as if you’d been crying,” he said.
“Oh, what a silly idea!” answered Letitia, with a laugh that would
have been quite successful on the stage, but could not deceive the
enamoured Tod; “I have a cold.”
“It’s not that you don’t look as pretty as usual. No matter what
you did, you’d always be out o’ sight. But it just gives me the willies
to think of your being down on your luck. Honest—I can’t stand it.”
Letitia looked away, more to avert her face from his searching
gaze than from embarrassment.
“Everybody gets blue now and then,” she said carelessly.
“But you oughtn’t to. I’m the one that ought to get blue—black
and blue.”
“I guess we all do, more or less.”
“If you’d just ease up on the way you keep giving me the marble
heart,” continued Tod, dropping his voice to the key of tenderness,
“I’d see to it that there’d never be a thing to make you blue.
Everything would go your way. I’d see to it.”
Letitia looked at him with a little vexed frown.
“Dear me, Tod!” she said crossly, “you’re not going to propose to
me here at dinner, are you, with everybody listening, too?”
Tod looked round rather guiltily. Letitia had exaggerated. The
only person who appeared to be noticing them was Mrs. Mortimer
Gault, and her glance immediately slipped away from his to give the
signal for withdrawal to a lady at the other end of the table.
IV
THE colonel’s visits now followed John Gault’s with businesslike
regularity. One week from the afternoon when the younger man
had paid his last call, Colonel Reed had made his customary
appearance and proffered his customary request.
With each succeeding gift of money his spirits seemed to rise, his
gracious bonhomie to become more pronounced. Upon this occasion
he had said cheerfully, as he dropped the pieces of gold into his old
chamois-skin purse:
“It’s these unconscionable tradespeople that eat up our
resources! Why can’t a provident government arrange things so that
we don’t have to pay butchers and bakers and milkmen? Life would
be so much better worth while if we could spend our money on
clothes and books and entertaining our friends than in paying bills.
Now, this”—jingling the gold in the purse—“goes to a son of Belial
who sells us groceries on tick.”
“Very kind of him, I should say,” said the other. “Aren’t you rather
lucky to have such good credit?”
“Well, that’s what I think,” said the colonel, throwing back his
head and laughing like an old prince in whom the joy of life and the
desire of the eyes still burned strong; “but Viola thinks credit is a
trap set by the king of all the devils.”
“Women are apt to be cautious about that sort of thing.”
“I don’t know about all women, but Viola is. She is more afraid of
credit than she is of smallpox. But I say to her: ‘My dear, look where
we would have been without it! And as long as these good,
charitable souls will give us food and drink for nothing, for goodness’
sake let them do it. Don’t let’s try and suppress such a worthy
impulse.’ Not, of course,” said the colonel, growing suddenly grave
and squaring his shoulders, “that we don’t intend to pay them. We
always do. Sometimes, it is true, we’re rather slow about it; but
eventually things are squared off to everybody’s satisfaction. How
else could we have the credit?”
He asked this question with an air of triumph that, to the listener,
seemed to have something in it of conscious cunning. Gault
answered with a commonplace about the advantage of inspiring so
great a trust in the vulgar mind. The colonel was openly gratified.
“Oh,” he said, as he moved toward the door, “there’s something
in the name of Ramsay Reed yet. But not enough,” he added,
laughing with a mischievous appreciation of the humor of his
misfortunes, “to let a grocery bill run on indefinitely. There was a
day when my name was good for any length of time—but that was
thirty years ago.”
Then he left, smiling and happy, and on the way home bought a
pot of pâté de foie gras, a bottle of claret, and a handkerchief
with an embroidered edge for Viola. At the grocery store on the
corner of the street where he lived he stopped and paid twenty
dollars on his bill, and then fared up the street with rapid strides, all
agog with pleasure at the thought of Viola’s delight in his present,
and the jolly little supper they would have on the end of the kitchen
table.
The man who had made these innocent pleasures possible was
far from enjoying those sensations of gratification said to be
experienced by a cheerful giver.
He had begun to know very dark hours. His first great love, come
tardily and reluctantly, at an age when the heart is almost closed to
soft influences and the mind is hardened with much worldly contact,
had come poisoned with torturing suspicions, with shame for his
own weakness, with fears of the truth.
Had he been a stronger man he would have torn up by the roots
this passion for a woman he dared not trust, have gone away and
tried to forget. But the lifelong habit of self-indulgence was too
powerful to be broken. He did not want to try and live without the
charm and torment of Viola’s presence. Had he been weaker he
would have yielded to the spell, never dared to question, and gone
on blindly into the purgatory of those who love and doubt. All his life
he had retained an ideal of womanhood—a creature aloof from the
coarseness of worldly ambition and vulgar greed. Now he found
himself bound to one the breath of whose life seemed to be tainted
with duplicity and sordid intrigue.
At times his state of uncertainty became intolerable. Then he
resolved to go to her, take her hands in his, and looking into her
eyes, ask for the truth. But the world’s lessons of a conventional
reserve, a well-bred reticence, asserted their claims, and he found
himself contemplating, with ironical bitterness, this picture of his
own simplicity. If they were deceiving him, how they would laugh—
laugh together—at the folly of the pigeon they were plucking so
cleverly! A life’s experience, caution, cynicism, had gone down into
dust before a girl’s gray eyes. Could she be false and those eyes look
into his so frankly and honestly? Could those lips, that folded on
each other in curves so full of innocence and truth, be ready with
words of hypocrisy and deceit? When he was with her such thoughts
seemed madness; when he was away from her his belief seemed a
miserable infatuation.
After the colonel’s last appearance he again determined to try
and see her alone. This, he discovered, was not as easy of
accomplishment as it had been on his first attempt. Arriving at the
house at four o’clock, he rang repeatedly, but was not able to gain
admittance. At last a small boy, who had been studying him through
the bars of the gate, volunteered the information that the lady was
out.
Gault turned away, and coming down the flagged walk, asked the
child if he knew what direction she had taken.
“I dunno that,” said the boy, “but she went out with her basket,
and when she goes with her basket she generally stays a long
while.”
Gault rewarded him for his information with a piece of money,
and turned down the street toward the other side of town.
It was a windy afternoon. The trades were just beginning, and
their clear, chill sweep had already borne away some of the evil
odors which hung about the old portion of the city. Gault could feel
the touch of fog in their buoyant breath, and knew that long tongues
of it like white wool were stealing in through the Golden Gate. The
city was putting on its summer aspect—a gray glare, softened by the
mingling of dust and haze that rode the breezes. Bits of paper, rags,
and straws were collected at corners in little whirling heaps.
Presently the mightier winds would come, winging their way across
miles of heaving seas to rush down the street in a mad carouse,
carrying before them the dirt and refuse and odors and uncleanness
which mark the dwelling of man.
He had walked some distance when, rounding a corner, a sharp
gust seized him. In its fierce exultation it threw a whirlwind of dust
into his eyes, so that, for a moment, he did not see that she was
coming toward him. Then he caught a glimpse of the approaching
figure and recognized it. She did not see him, but was engaged in
her customary amusement of looking into the gardens. There was
an air of unmistakable alertness and gaiety about her. Her hand
tapped the tops of the fence-rails as she came, and she looked at
the floral display behind them with happy eyes. Her scanty black
skirt was sometimes whirled round her feet, showing her small
ankles and narrow russet shoes. Once she had to put up her hand to
her hat,—a white sailor bound with a dark ribbon,—and the
frolicsome wind swept all the loosened ends of her hair forward and
lashed her skirts out on either side. She had a basket on one arm,
and holding this firmly, leaned back almost on the wind, laughing to
herself.
At the same moment she caught sight of him. The wind dropped
suddenly, as if conscious that she should not be presented in such
boisterous guise to a lover’s eye, and her figure seemed to fall back
into lines of decorous demureness; only the color and laughter of
her recent buffeting still remained in her face.
“Is it you?” she cried. “Did you see me in the wind? Isn’t it fun?”
They met, and he took her hand. She was all blown about, but
fresh as a flower that has shaken off the dew. The contrast between
them, between what might be called their different ranks in society,
was much more clearly marked in the open light of the street than in
the ragged homeliness of her own parlor.
While he was essentially the man of luxurious environment and
assured position, she presented the appearance of a working-girl.
Even the delicacy and refinement of her face could not counteract
the suggestion of her dress. Beauty when unadorned may adorn the
most, but it cannot give to ill-made old clothes the effect of
garments made by a French modiste. John Gault was used to
women who wore this kind of clothes—so used, in fact, that he
hardly knew what made Viola appear so different from the other
girls of his acquaintance. The contrast in their looks seemed to mark
more clearly the contrast in their positions, seemed to purposely
accentuate that wide gulf set between them.
Gault took her basket from her and dropped into place at her
side. The high rows of houses protected them from the wind, and
only as they crossed the open spaces at the intersection of streets
did it catch them, and, for a moment, play boisterously with them.
The girl seemed in excellent spirits. He had noticed this with
every recurring visit. Looking back upon her as she was when he
had first known her, care-worn, pale, and quiet, she seemed now
like a different person. Her glance sparkled with animation, her voice
was full of that thrilling quality which some women’s voices acquire
in moments of happiness. She was a hundred times more fatally
alluring than she had been in the beginning. He knew now that while
he was with her his reason would always be in abeyance to his
heart.
“You seem to be in very good spirits,” he said to her, not without
a feeling of personal grievance that some cause of which he was
ignorant should add so to her lightness of heart.
“I am,” she answered. “I’m in very good spirits. I’m quite happy.
It’s something lovely to feel so gay in your heart, isn’t it?”
“I don’t know; maybe I’ve never felt so.”
“Oh, what nonsense!” she cried, looking at him reproachfully.
“You, who have always had just what you wanted! I used to be
afraid of you at first. It seemed rather awful to know anybody who’d
always had things go exactly their way.”
He ignored the remark and said:
“What’s making you happy? Tell it to me, and then perhaps I’ll
get a little reflection of it.”
“I don’t know that it’s any one especial thing. Happiness comes
when lots of little things fit nicely together. I never had one big thing
in a lump to make me happy. I tell you what’s doing a good deal
toward it. Father and I are”—she made an instant’s pause and then
said—“doing so much better; financially, I mean. It’s such a relief!
You don’t know.”
He turned and looked at her and met her eyes. They looked
rather abashed, and then fell away from the scrutiny of his.
“You don’t think it queer of me to tell you that, do you?” she
asked. “I tell you a good many things I wouldn’t say to other
people.”
“I am proud that you should have such confidence in me.”
“Well,” she continued, with a quick sigh of relief, “we’ve been
lately—that is, just about when we first knew you, and before that—
really quite badly off. And my father being so sanguine, and having
once been so differently situated, it’s very hard on him—very hard.”
She paused, and he felt that she was looking at him for
confirmation of her remark.
“Very; I quite understand,” he answered.
“And, really, it was dreadful. It’s trying for old people—so much
anxiety. And then, just at the very worst, things suddenly
brightened. Just about a month or six weeks ago the luck changed.
You must have been the mascot.”
This time he looked at her, but her glance was averted.
“Go on,” he said, thinking that his voice sounded strange.
“Because it was after we knew you that things began to get
better. I was angry with my father that first day when he asked you
in, because I didn’t want you to see how—how straitened we were.
There’s a pride of poverty, you know; well, I suppose I must have a
little bit of it. Everything was at its worst then. But now it’s all
different. You’ve been the mascot.”
He again felt her eyes surveying him, but found it impossible to
look at her. In his heart he was afraid of what he might read in her
face.
“Don’t you like being a mascot?” she queried, in her happy girl’s
voice. “You don’t look as if you did.”
“I’m proud and flattered, probably too much so for speech.”
“I’m glad, because that’s what you were. There’s no getting out
of it. I’ll tell you how it happened. My father used to own a great
deal of stock in mines and companies and things, and when
everything went down so fast, he sold almost all of it. But some he
kept. He had it put away in the drawers of his desk up-stairs in his
room, and about two months ago it began to go up, and now it pays
dividends and we get them. Isn’t that good luck?”
She was close to him, looking into his face. He turned his head
this time and confronted her with a steady gaze. In the harsh
afternoon light every curve and line of her countenance was
revealed. Her eyes were full of light and joy. His glance met and held
them for one searching moment, then turned away baffled.
“Very good luck. I congratulate you,” he said.
“You may well,” she answered. “I’d given up expecting good luck
ever any more in this world. I believe in it, and my father’s had come
and gone almost before I was born, and mine—mine hasn’t come
yet, I suppose.”
“Unless you discover some more old stock in the pigeonholes of
the desk.”
“Oh, I don’t think that’s likely. Lightning doesn’t strike twice in
the same place.”
“What was the stock? Mining stock?”
She seemed in doubt for a moment, then said:
“Yes, I think so—yes, surely, mining stock.”
“Do you remember the name of the mine?”
He glanced at her as she walked beside him. She appeared to be
cogitating.
“I don’t believe I do,” she answered at length. “To tell you the
truth, I don’t believe my father mentioned it to me. I’m very stupid
about business. I’ve never had any necessity to know about it, and
so I’ve never learned.”
“How long had it been lying in the desk?”
“Oh, years and years! Probably twenty. It was a relic of the days
when everything was booming.”
“If he’s been paying assessments on it all these years, he ought
certainly to be repaid now.”
He was scrutinizing her sharply. Her profile was toward him, and
at this remark he saw the color mount into her cheek, and that
curious appearance of immobility come over her face which denotes
a sudden, almost electric stoppage and then concentration of mental
activity. She raised her head and said, without looking at him:
“Assessments are a yearly or semi-yearly payment, aren’t they?”
“Yes, or quarterly—according to the way the stock is drawn.”
“But isn’t there some that is non-assessable? I’ve surely heard
that expression.”
“In other States, but in California—well, possibly there might be.”
“I’m sure there must be. This of my father’s must have been.”
She came quite close to him in her earnestness, and looked at him
with an expression of uneasiness on her face.
“It must have been that kind,” she insisted; “probably you never
heard of this mine.”
“Probably I never did,” he answered grimly.
They walked on for a few moments in silence. There was a visible
drop in her spirits. Stealing a side glance at her, he could see that
she was looking down, evidently in troubled thought. Suddenly she
raised her head and said:
“Well, I don’t really know anything about it. Only I do hope one
thing, and that is that it will go on paying.”
“Don’t bother about that,” he said; “it will.”
“What makes you think it will?”
He turned on her roughly and said:
“Don’t you think it will?”
“I’d like to think so,” she answered, abashed by his unusual
manner; “but I’ve learned that it’s foolish to hope. I try not to.”
He gave a short, disagreeable laugh and said:
“Oh, not in this case. Hope as much as you like.”
“You’re very cheering,” she answered; “but I don’t see how you
can be so sure.”
“It seems to me you’re very pessimistic—especially for a young
woman who has just found a drawerful of paying stock.”
His manner in making this remark was so impregnated with
angry bitterness that Viola, chilled and repelled, made no response.
In silence they walked onward till a turn in the street brought them
in sight of the house.
At the gate she said rather timidly:
“Would you like to come in?”
He had been carrying the basket, and now found the depositing
of it in a place of safety an excuse to enter; for even in his present
state of morose ill humor he could not forego the pleasure of a few
more moments of her society.
In the cold, half-furnished house their footsteps echoed with a
strangely solitary effect. She preceded him into the parlor, and
moved about with the confident tread of the chatelaine, pulling up
the blinds, putting the basket out of sight, and laying aside her hat
and gloves. There were some thin flowered muslin curtains hanging
over the bay-window, and she arranged the folds of these with deft,
proprietary touches, and then stepped back and studied the effect.
After watching her for a moment the visitor said in a tone of
restored amiability:
“Aren’t those something new?”
She looked at him with quick, grateful recognition of his change
of mood.
“Yes; do you like them? I changed my mind about a dozen times
before I bought them. Even now I don’t know whether I’m entirely
satisfied.”
“Oh, you ought to be,” he said, as he drew near and eyed the
curtains with the air of a connoisseur; “I’m sure you couldn’t have
chosen anything prettier.”
Viola’s spirits rose to the level they had been at when he met her
earlier in the afternoon. Her eyes brightened and her face took on its
most animated expression.
“They’re another outward and visible sign of the rise in mining
stock,” she continued. “I’m so glad you noticed them without my
having to make you do so.”
“Do you want to know why I did?”
“Because they were pretty, of course.”
“Not at all. I was looking at you as you arranged them, and
wondering why a pair of curtains should be so much more
interesting than I was.”
“What made you think they were?”
“Because you were devoting yourself to them and coldly ignoring
me.”
“That was because I was a little bit frightened of you. You were
so cross just now, before we came in, that I didn’t know what to say
to you.”
“I cross? What a calumny! I was in my sweetest humor.”
She looked at him mischievously.
“If you call that your sweetest humor, all I can think is that you’re
not as clever as you pretend to be.”
“I’m afraid I’m not. For example, I’m not clever enough to
understand you—a little girl like you, scarcely half my age.”
“Am I really such a sphinx?”
“You are to me.”
“I like that,” she said, smiling, and gathering up the edge of the
curtain in a frill; “I don’t want everybody to see through me. But
you’re different.”
“How am I different?”
“You’re more a friend than other people—more a friend than
anybody else I know. Tell me what you don’t understand about me,
and I’ll explain it. I won’t leave myself a single secret.”
Though he was standing close to her, looking down at her, he
suddenly dropped his voice to the key that was the lowest she could
hear.
“If I only dared to ask, and you would only tell the truth.”
“Dared to ask!” she repeated blankly, alarmed and upset by his
singular change of manner.
“And you would tell the truth,” he added, and heard his own
voice sound suddenly husky and shaken. “Tell it to me now!”
“I always do,” she stammered.
“No matter what it is,” he continued, as if he had not heard her
—“no matter how it may hurt me or injure you.”
The color ran over her face and as quickly ebbed away, leaving
her pallid. It might have been the confession of innocence or the
confusion of guilt. She looked nervously from side to side, raised her
eyes to his, and dropped them again.
“There are always a few things a person can’t tell,” she almost
whispered.
He gave an ugly laugh, and put his arm half round her as if to
draw her to him, then drew back as quickly, and turning away,
walked to the window. Viola did not seem to have noticed the
attempted caress. There was a moment of penetrating silence. He
wondered if she could hear his heart beat.
Then she said:
“Why do you say such strange things? I always tell you the
truth.”
To his listening ear her voice sounded affectedly naïve. He
answered without moving:
“Of course you do. So do all women since the days of Eve.”
“But you don’t seem to believe me.”
“You mustn’t jump at such hasty conclusions.”
“Have you heard anything about me that would make you think I
was deceitful?”
“I have never spoken of you to any one except your father.”
“I can’t understand you at all to-day. You’re so changeable and
moody, and sometimes so ill-humored.”
“What a dreadful afternoon you’ve had! I’m sorry.” Then, with an
abrupt change of tone: “Who picks up the leaves of the deodar and
ties them up in those neat little bundles?”
“I do—do you believe me?” She spoke with a sharpness he had
never heard her use before.
He broke out into sudden laughter that this time sounded
genuine. Turning from the window, he came toward her and took
her hand.
“Are you angry?” he asked. “I don’t wonder. Say the most
disagreeable things you can think of, and they won’t be more than I
deserve.”
For the second time this afternoon she beamed over his
restoration to good humor.
“I’m not a very good person to quarrel with,” she said, looking at
him with soft, forgiving eyes, “though, as you see, I’ve got a
temper.”
He gave her hand a little pressure and relinquished it, taking up
his hat.
“Accept a hundred apologies from me for my rudeness. Good-by.”
“You were disagreeable,” she admitted, as they went together
into the hall. “You seemed as if you didn’t believe half I said to you,
and actually as if our good luck made you angry.”
Gault had opened the door, and his face was turned from her.
“Oh, don’t think that,” he answered, as he stepped out on to the
porch; “whatever gives you happiness adds to mine. Adios,
Welcome to Our Bookstore - The Ultimate Destination for Book Lovers
Are you passionate about books and eager to explore new worlds of
knowledge? At our website, we offer a vast collection of books that
cater to every interest and age group. From classic literature to
specialized publications, self-help books, and children’s stories, we
have it all! Each book is a gateway to new adventures, helping you
expand your knowledge and nourish your soul
Experience Convenient and Enjoyable Book Shopping Our website is more
than just an online bookstore—it’s a bridge connecting readers to the
timeless values of culture and wisdom. With a sleek and user-friendly
interface and a smart search system, you can find your favorite books
quickly and easily. Enjoy special promotions, fast home delivery, and
a seamless shopping experience that saves you time and enhances your
love for reading.
Let us accompany you on the journey of exploring knowledge and
personal growth!
ebookgate.com