Table of Contents

The One-Dimensional Particle in a Box

Cyanine Dyes

🥅 Learning Objectives

Cyanine Dyes

Cyanine Dyes Cyanine dye molecules are often modelled as one-dimension particles in a box. To understand why, start by thinking classically. You learn in organic chemistry that electrons can more “freely” along alternating double bonds. If this is true, then you can imagine that the electrons can more from one Nitrogen to the other, almost without resistance. On the other hand, there are sp3-hybridized functional groups attached to the Nitrogen atom, so once the electron gets to Nitrogen atom, it has to turn around and go back whence it came. A very, very, very simple model would be to imagine that the electron is totally free between the Nitrogen atoms, and totally forbidden from going much beyond the Nitrogen atoms. This suggests modeling these systems a potential energy function like: V(x)={+x000<x<a+ax where a is the length of the box. A reasonable approximate formula for a is a=(5.67+2.49(k+1))1010 m

Postulate: The squared magnitude of the wavefunction is proportional to probability

What is the interpretation of the wavefunction? The Born postulate indicates that the squared magnitude of the wavefunction is proportional to the probability of observing the system at that location. E.g., if ψ(x) is the wavefunction for an electron as a function of x, then p(x)=|ψ(x)|2 is the probability of observing an electron at the point x. This is called the Born Postulate.

The Wavefunctions of the Particle in a Box (boundary conditions)

Particle in a Box Potential

The nice thing about this “particle in a box” model is that it is easy to solve the time-independent Schrödinger equation in this case. Because there is no chance that the particle could ever “escape” an infinite box like this (such an electron would have infinite potential energy!), |ψ(x)|2 must equal zero outside the box. Therefore the wavefunction can only be nonzero inside the box. In addition, the wavefunction should be zero at the edges of the box, because otherwise the wavefunction will not be continuous. So we should have a wavefunction like ψ(x)={0x0?????0<x<a0ax

Postulate: The wavefunction of a system is determined by solving the Schrödinger equation

How do we find the wavefunction for the particle-in-a-box or, for that matter, any other system? The wavefunction can be determined by solving the time-independent (when the potential is time-independent) or time-dependent (when the potential is time-dependent) Schrödinger equation.

The Wavefunctions of the Particle in a Box (solution)

Wavefunction gif To find the wavefunctions for a system, one solves the Schrödinger equation. For a particle of mass m in a one-dimensional box, the (time-independent) Schrödinger equation is: (22md2dx2+V(x))ψn(x)=Enψn(x) where V(x)={+x000<x<a+ax We already deduced that ψ(x)=0 except when the electron is inside the box (0<x<a), so we only need to consider the Schrödinger equation inside the box: (22md2dx2)ψn(x)=Enψn(x) There are systematic ways to solve this equation, but let's solve it by inspection. That is, we need to know:

Question: What function(s), when differentiated twice, are proportional to themselves?\

This suggests that the eigenfunctions of the 1-dimensional particle-in-a-box must be some linear combination of sine and cosine functions, ψn(x)=Asin(cx)+Bcos(dx) We know that the wavefunction must be zero at the edges of the box, ψ(0)=0 and ψ(a)=0. These are called the boundary conditions for the problem. Examining the first boundary condition, 0=ψ(0)=Asin(0)+Bcos(0)=0+B indicates that B=0. The second boundary condition 0=ψ(a)=Asin(ca) requires us to recall that sin(x)=0 whenever x is an integer multiple of π. So c=nπ where n=1,2,3,. The wavefunction for the particle in a box is thus, ψn(x)=Ansin(nπxa)n=1,2,3,

Normalization of Wavefunctions

As seen in the previous section, if a wavefunction solves the Schrödinger equation, any constant multiple of the wavefunction also solves the Schrödinger equation, H^ψ(x)=Eψ(x)H^(Aψ(x))=E(Aψ(x)) Owing to the Born postulate, the complex square of the wavefunction can be interpreted as probability. Since the probability of a particle being at some point in space is one, we can define the normalization constant, A, for the wavefunction through the requirement that: |ψ(x)|2dx=1. In the case of a particle in a box, this is: (1)1=|ψn(x)|2dx(2)=0aψn(x)ψn(x)dx(3)=0aAnsin(nπxa)(Ansin(nπxa))dx(4)=|An|20asin2(nπxa)dx To evaluate this integral, it is useful to remember some trigonometric identities. (You can learn more about how I remember trigonometric identities here.) The specific identity we need here is sin2x=12(1cos2x):

(5)1=|An|20asin2(nπxa)dx(6)=|An|20a12(1cos(2nπxa))dx(7)=|An|22(0a1dx0acos(2nπxa)dx)(8)=|An|22([x]0a[a2nπsin(2nπxa)]0a)(9)=|An|22(a0)

So |An|2=2a Note that this does not completely determine An. For example, any of the following normalization constants are allowed, An=2a=2a=i2a=i2a In general, any square root of unity can be used, An=(k±i1k2)2a where k is any real number. The arbitrariness of the phase of the wavefunction is an important feature. Because the wavefunction can be imaginary (e.g., if you choose An=i2a), it is obvious that the wavefunction is not an observable property of a system. The wavefunction is only a mathematical tool for quantum mechanics; it is not a physical object.

Summarizing, the (normalized) wavefunction for a particle with mass m confined to a one-dimensional box with length a can be written as: ψn(x)=2asin(nπxa)n=1,2,3,

Note that in this case, the normalization condition is the same for all n; that is an unusual property of the particle-in-a-box wavefunction.

While this normalization convention is used 99% of the time, there are some cases where it is more convenient to make a different choice for the amplitude of the wavefunctions. I say this to remind you that normalization the wavefunction is something we do for convenience; it is not required by physics!

Normalization Check

One advantage of using Jupyter is that we can easily check our (symbolic) mathematics. Let's confirm that the wavefunction is normalized by evaluating, 0a|ψn(x)|2dx

The Energies of the Particle in a Box

How do we compute the energy of a particle in a box? All we need to do is substitute the eigenfunctions of the Hamiltonian, ψn(x) back into the Schrödinger equation to determine the eigenenergies, En. That is, from H^ψn(x)=Enψn(x) we deduce (10)22md2dx2(Ansin(nπxa))=En(Ansin(nπxa))(11)An22mddx(nπacos(nπxa))=En(Ansin(nπxa))(12)An22m(nπa)2sin(nπxa)=En(Ansin(nπxa))(13)2n2π22ma2=En Using the definition of , we can rearrange this to: (14)En=2n2π22ma2n=1,2,3,(15)=h2n28ma2 Notice that only certain energies are allowed. This is a fundamental principle of quantum mechanics, and it is related to the "waviness" of particles. Certain "frequencies" are resonant, and other "frequencies" cannot be observed. The only energies that can be observed for a particle-in-a-box are the ones given by the above formula.

Zero-Point Energy

Naïvely, you might expect that the lowest-energy state of a particle in a box has zero energy. (The potential in the box is zero, after all, so shouldn't the lowest-energy state be the state with zero kinetic energy? And if the kinetic energy were zero and the potential energy were zero, then the total energy would be zero.)

But this doesn't happen. It turns out that you can never "stop" a quantum particle; it always has a zero-point motion, typically a resonant oscillation about the lowest-potential-energy location(s). Indeed, the more you try to confine a particle to stop it, the bigger its kinetic energy becomes. This is clear in the particle-in-a-box, which has only kinetic energy. There the (kinetic) energy increases rapidly, as a2, as the box becomes smaller: Tn=En=h2n28ma2 The residual energy in the electronic ground state is called the zero-point energy, Ezero-point energy=h28ma2 The existence of the zero-point energy, and the fact that zero-point kinetic energy is always positive, is a general feature of quantum mechanics.

Zero-Point Energy Principle: Let V(x) be a nonnegative potential. The ground-state energy is always greater than zero.

More generally, for any potential that is bound from below, Vmin=minxV(x) the ground-state energy of the system satisfies Ezero-point energy>Vmin.

Nuance: There is a tiny mathematical footnote here; there are some V(x) for which there are no bound states. In such cases, e.g., V(x)=constant, it is possible for E=Vmin.)

Atomic Units

Because Planck's constant and the electron mass are tiny numbers, it is often useful to use atomic units when performing calculations. We'll learn more about atomic units later but, for now, we only need to know that, in atomic units, , the mass of the electron, me, the charge of the electron, e, and the average (mean) distance of an electron from the nucleus in the Hydrogen atom, a0, are all defined to be equal to 1.0 in atomic units. (16)=h2π=1.0545718×1034 J s=1 a.u.(17)me=9.10938356×1031 kg=1 a.u.(18)e=1.602176634×1019 C=1 a.u.(19)a0=5.291772109×1011 m=1 Bohr=1 a.u. The unit of energy in atomic units is called the Hartree, Eh=4.359744722×1018 J=1 Hartree=1 a.u. and the ground-state (zero-point) energy of the Hydrogen atom is 12Eh.

We can now define functions for the eigenenergies of the 1-dimensional particle in a box:

Postulate: The wavefunction contains all the physically meaningful information about a system.

While the wavefunction is not itself observable, all observable properties of a system can be determined from the wavefunction. However, just because the wavefunction encapsulates all the observable properties of a system does not mean that it contains all information about a system. In quantum mechanics, some things are not observable. Consider that for the ground (n=1) state of the particle in a box, the root-mean-square average momentum, p¯rms=2mT=(2m)h2n28ma2=hn2a increases as you squeeze the box. That is, the more you try to constrain the particle in space, the faster it moves. You can't "stop" the particle no matter how hard you squeeze it, so it's impossible to exactly know where the particle is located. You can only determine its average position.

Postulate: Observable Quantities Correspond to Linear, Hermitian Operators.

The correspondence principle says that for every classical observable there is a linear, Hermitian, operator that allows computation of the quantum-mechanical observable. An operator, C^ is linear if for any complex numbers a and b, and any wavefunctions ψ1(x) and ψ2(x), C^(aψ1(x,t)+bψ2(x,t))=aC^ψ1(x,t)+bC^ψ2(x,t) Similarly, an operator is Hermitian if it satisfies the relation, ψ1(x,t)C^ψ2(x,t)dx=(C^ψ1(x,t))ψ2(x,t)dx or, equivalently, ψ1(x,t)(C^ψ2(x,t))dx=ψ2(x,t)(C^ψ1(x,t))dx That is for a linear operator, the linear operator applied to a sum of wavefunctions is equal to the sum of the linear operators directly applied to the wavefunctions separately, and the linear operator applied to a constant times a wavefunction is the constant times the linear operator applied directly to the wavefunction. A Hermitian operator can apply forward (towards ψ2(x,t)) or backwards (towards ψ1(x,t)). This is very useful, because sometimes it is much easier to apply an operator in one direction.

We've already been introduced to the quantum-mechanical operators for the momentum, p^=iddx and the kinetic energy, T^=22md2dx2 These operators are linear because the derivative of a sum is the sum of the derivatives, and the derivative of a constant times a function is that constant times the derivative of the function. These operators are also Hermitian. For example, to show that the momentum operator is Hermitian:

(20)ψ1(x,t)p^ψ2(x,t)dx=ψ1(x,t)(iddx)ψ2(x,t)dx(21)=iddx(ψ1(x,t)ψ2(x,t))(ψ2(x,t)ddxψ1(x,t))dx

Here we used the product rule for derivatives, f(x)dgdx=df(x)g(x)dxg(x)dfdx. Using the fundamental theorem of calculus and the fact the probability of observing a particle at ± is zero, and therefore the wavefunctions at infinity are also zero, one knows that ddx(ψ1(x,t)ψ2(x,t))=[ψ1(x,t)ψ2(x,t)]=0 Therefore the above equation can be simplified to (22)ψ1(x,t)p^ψ2(x,t)dx=iψ2(x,t)ddxψ1(x,t)dx(23)=ψ2(x,t)iddxψ1(x,t)dx(24)=ψ2(x,t)(iddxψ1(x,t))dx(25)=ψ2(x,t)(p^ψ1(x,t))dx

The expectation value of the momentum of a particle-in-a-box is always zero. This is intuitive, since electrons (on average) are neither moving to the right nor to the left inside the box: if they were, then the box itself would need to be moving. Indeed, for any real wavefunction, the average momentum is always zero. This follows directly from the previous derivation with ψ1(x,t)=ψ2(x,t). Thus: (26)ψ2(x,t)p^ψ2(x,t)dx=iψ2(x,t)ddxψ2(x,t)dx(27)=iψ2(x,t)(ddxψ2(x,t))dx(28)=0 The last line follows because the only number that is equal to its negative is zero. (That is, x=x if and only if x=0.) It is a subtle feature that the eigenfunctions of a real-valued Hamiltonian operator can always be chosen to be real-valued themselves, so their average momentum is clearly zero. We often denote quantum-mechanical expectation values with the shorthand, p^=0

The momentum-squared of the particle-in-a-box is easily computed from the kinetic energy, p^2=0aψn(x)p^2ψn(x)dx=0aψn(x)(2mT^)ψn(x)dx=2mEn=h2n24a2

Intuitively, since the box is symmetric about x=a2, the particle has equal probability of being in the first-half and the second-half of the box. So the average position is expected to be x=a2

We can confirm this by explicit integration,

(29)x=0aψn(x)xψn(x)dx(30)=0a(2asin(nπxa))x(2asin(nπxa))dx(31)=2a0axsin2(nπxa)dx(32)=2a[x24xsin2nπxa4nπacos2nπxa8n2π2a2]0a(33)=2a[a2400](34)=a2

Similarly, we expect that the expectation value of x2 will be proportional to a2. We can confirm this by explicit integration, \begin{align} \langle x^2 \rangle &= \int_0^a \psi_n^*(x)\, x^2 \,\psi_n(x) dx \ &= \int_0^a \left(\sqrt{\tfrac{2}{a}} \sin\left(\tfrac{n \pi x}{a} \right)\right) x^2 \left(\sqrt{\tfrac{2}{a}}\sin\left(\tfrac{n \pi x}{a} \right)\right) dx \ &= \tfrac{2}{a} \int_0^a x^2 \sin^2\left(\tfrac{n \pi x}{a} \right) dx \ &= \tfrac{2}{a} \left[ \tfrac{x^3}{6}

We can verify these formulas by explicit integration.

Heisenberg Uncertainty Principle

The previous example gives a first example of the more general Heisenberg Uncertainty Principle. One specific manifestation of the Heisenberg Uncertaity Principle is that the variance of the position, σx2=x2x2, times the variance of the momentum. σp2=p^2p^2 is greater than 24. We can verify this formula for the particle in a box. (35)24σx2σp2(36)=(a2[1312n2π2]a24)(h2n24a20)(37)=(a2[1312n2π2]a24)(2π2n2a2)(38)=2π2n2(11212n2π2) The right-hand-side gets larger and larger as n increases, so the largest value occurs where: (39)242π2(11212π2)(40)=0.322472

Double-Checking the Energy of a Particle-in-a-Box

To check the energy of the particle in a box, we can compute the kinetic energy density, then integrate it over all space. That is, we define: τn(x)=ψn(x)(22md2dx2)ψn(x) and then the kinetic energy (which is the energy for the particle in a box) is Tn=0aτn(x)dx

Note: In fact there are several different equivalent definitions of the kinetic energy density, but this is not very important in an introductory quantum chemistry course. All of the kinetic energy densities give the same total kinetic energy. However, because the kinetic energy density, τ(x), represents the kinetic energy of a particle at the point x, and it is impossible to know the momentum (or the momentum-squared, ergo the kinetic energy) exactly at a given point in space according to the Heisenberg uncertainty principle), there can be no unique definition for τ(x).

Visualizing the Particle-in-a-Box Wavefunctions, Probabilities, etc.

In the next code block, the wavefunction, probability density, derivative, second-derivative, and kinetic energy density for the particle-in-a-box are shown. Notice that the kinetic-energy-density is proportional to the probability density, and that the first and second derivatives are not zero at the edge of the box, but the wavefunction and probability density are. It's useful to change the parameters in the below figures to build your intuition for the particle-in-a-box.

🪞 Self-Reflection

🤔 Thought-Provoking Questions

🔁 Recapitulation

🔮 Next Up...

📚 References

My favorite sources for this material are:

There are also some excellent wikipedia articles: