Ideas of Quantum Chemistry P19 potx

10 268 0
Ideas of Quantum Chemistry P19 potx

Đang tải... (xem toàn văn)

Thông tin tài liệu

146 4. Exact Solutions – Our Beacons The solution to this equation is given by (4.1), which may also be written as  =A sinκx +B cosκx (4.3) with κ 2 = 2mE ¯ h 2  (4.4) Now, the key is to recall (p. 74, Fig. 2.5), that the wave function has to be con- tinuous and, therefore, two conditions have to be fulfilled: 1)  =0forx =0and 2)  =0forx =L. The first condition immediately gives B =0, the second in this situation is equivalent to κL =nπ,forn =01. From this follows energy quanti- zation, because κ contains energy E. One obtains, therefore, the following solution (a standing wave 7 ): E n = n 2 h 2 8mL 2 n=1 2 3 (4.5)  n =  2 L sin nπ L x n =1 2 3 (4.6) because n = 0 has to be excluded as leading to the wave function equal to zero everywhere, while n<0 may be safely excluded as leading to the same wave func- tions as 8 n>0. Fig. 4.2 shows the wave functions for n =1, 2, 3. 2D rectangular box Let us consider a rectangular box (Fig. 4.3) with sides L 1 and L 2 and V =0inside and V =∞outside. We very easily obtain the solution to the Schrödinger equa- tion after a straightforward separation of variables x and y leading to the two 1D Schrödinger equations. The energy eigenvalue is equal to the sum of the energies for the 1D problems E n = h 2 8m  n 2 1 L 2 1 + n 2 2 L 2 2   (4.7) while the wave function has form of the product  n 1 n 2 =2  1 L 1 L 2 sin n 1 π L 1 x ·sin n 2 π L 2 y (4.8) where n 1 n 2 =1 2 7 Recall that any stationary state has a trivial time-dependence through the factor exp(−i E ¯ h t).Astand- ing wave at any time t has a standing-still pattern of the nodes i.e. the points x with  =0. 8 With the opposite sign, but it does not matter. 4.2 Particle in a box 147 Fig. 4.2. The wave functions for the particle in a box corresponding to n = 1, 2, 3. Note the increasing number of nodes, when the energy E i of the sta- tionary state increases. Example 1. Butadiene naively The particle-in-box problem has more to do with chemistry than would appear at first glance. In organic chemistry, we consider some molecules with conjugate double and single bonds, one of the simplest is butadiene: =−= What does this molecule have to do with the particle in a box? It seems nothing. First, we have not a single particle but 40 particles (10 nuclei and 30 electrons), second, where is this constant potential for the motion of the particle? Nowhere. Third, a molecule does not represent a one-dimensional but a three-dimensional object, and in addition, a curved one instead of a beautiful section of the x axis. It would seem that any attempt to apply such a primitive theory to our molecule is ridiculous and yet in such a difficult situation we will see the power of the exact so- lutions reported in the present chapter. All above objections are perfectly justified, but let us try to simplify our system a little. In the molecule under study the CC bonds are “averaged”, which facilitates the motion of the π electrons along the system (this notion will become clear in Chapter 8; the π electrons are loosely bound to the molecule, we may assume that other electrons are always rigidly bound and will therefore be ignored). If • we removed the π electrons from the molecule (and put them temporarily into a safe), and then • “ground up” the remaining (positively charged) molecular core and distributed the ground mass uniformly along the x axis within a section of length L equal to 148 4. Exact Solutions – Our Beacons Fig. 4.3. Examples of the wave functions for a particle in a square box, the quantum numbers (n 1 n 2 ) correspond to: a) (1 1);b)(1 2);c)(2 1);d)(2 2);e)(4 4). The background colour corresponds to zero. In the case shown the higher the energy the more nodes in the wave function. This rule is not generally true. For example, in a rectangular box with L 1  L 2 even a large increase of n 1 does not raise the energy too much, while introducing a lot of nodes. On the other hand, increasing n 2 by 1 raises the energy much more, while introducing only one extra node. A reader acquainted with hydrogen atom orbitals will easily recognize the resemblance of the above figures to some of them (cf. pp. 180–185), because of the rule mentioned above. the length of the molecule (averaging the potential energy for a charged particle) to construct a highway for the π electrons • added the first π electron from the safe, then 4.2 Particle in a box 149 this single electron would represent something similar to a particle in a box. 9 As- suming this simplified model we know all the details of the electron distribution, including the ground-state and excited-state wave functions (in the one-particle case called the orbitals). If we now took all the π electrons from the safe, added them one by one to the system, assuming that they would not see one another, 10 then taking into account the Pauli exclusion principle (described in more detail in Chapter 8) we would obtain information about the electron density distribution in the molecule. The idea we are describing is called the Free Electron Molecular Orbitals (FEMO) method. FEMO method In our example, the total electron density distribution (normalized to four π electrons, i.e. giving 4 after integration over x)isgivenas 11 ρ(x) =2ψ 2 1 +2ψ 2 2 =2 2 L sin 2 π L x +2 2 L sin 2 2π L x = 4 L  sin 2 π L x +sin 2 2π L x   The function ρ(x) is shown in Fig. 4.4.a. It is seen that: 1. ρ(x) is the largest on the outermost bonds in the molecule, exactly where chemists put their two little lines to symbolize a double bond. 2. π-electron density, i.e. ρ(x) is non-zero in the centre. This means that the bond over there is not strictly a single bond. This key information about the butadiene molecule has been obtained at prac- tically no cost from the simple FEMO model. Of course, we cannot expect the description to reflect all the details of the charge distribution in the butadiene molecule, but one may expect this approach to be able to reflect at least some rough features of the π electron distribution. If the results of more advanced calculations contradict the rough particle-in-box results, then we should take a closer look at them and search for an error. This is the strength of the simple exact model systems. They play the role of the beacons – points of reference. 4.2.2 CYCLIC BOX The 1D box described above is similar to a stick in which the particle can move. The butadiene molecule is rather similar to such a stick and, therefore, the 1D box models it quite well. 9 Almost, because the potential is not quite constant (ends!). Also one might remove the particle from the box at the expense of a large but finite energy (ionization), which is not feasible for the particle in a box. 10 As we will see in Chapter 8, this approximation is more realistic than it sounds. 11 The student “i” is characterized by a probability density distribution ρ i (x) of finding him at co- ordinate x (we limit ourselves to a single variable, measuring his position, say, on his way from the dormitory to the university). If all students moved independently, the sum of their individual proba- bility densities at point x 0 ,i.e.ρ(x 0 ) =  i ρ i (x 0 ) would be proportional to the probability density of finding any student at x 0  The same pertains to electrons, when assumed to be independent. 150 4. Exact Solutions – Our Beacons Fig. 4.4. π-electron density charge distributions for several molecules computed by the FEMO method. The length of each molecule L has been assumed to be equal 1. For other lengths the charge distribu- tions are similar. The electron density for four electrons in butadiene (a) and of six electrons in hexa- triene (b). The electron density maxima coincide with the positions chemists write as double bonds. The six electron density distribution in the benzene molecule is peculiar, because it is constant along the perimeter of the molecule (c). If we subtract an electron from benzene (d) or add an electron to it (e), then maxima and minima of the π electron density appear. If an electron is subtracted (d) there are two maxima (double bonds) and two π electron deficient regions denoted as having charge + 1 2 .Afterone electron π is added (e) then we obtain four maxima (two double bonds and two electron-rich regions denoted by charge − 1 2 ). 4.2 Particle in a box 151 And what can model the benzene molecule? In a crude approximation we may think of benzene as a stick with the two ends joined in such a way as to be unable to recognize where the union has taken place. Limiting ourselves to this effect, 12 we may use the solution given by (4.3) and impose appropriate boundary condi- tions. What could these boundary conditions be? The wave function at the two ends of the box has to be stitched together without leaving any trace of the seam. This is achieved by two boundary conditions: (0) =(L) forcing the two wave function values to match and   (0) =  (L) making the seam “invisible”. The two conditions mean: A sinκ0 +B cosκ0 =A sin κL +B cos κL Aκ cosκ0 −Bκsinκ0 = AκcosκL −BκsinκL or B =A sinκL +B cosκL A =AcosκL −B sinκL To find a non-trivial solution the determinant of the coefficients at the unknown quantities A and B has to vanish: 13     sinκL cosκL −1 cosκL −1 −sinκL     =0 which is equivalent to cosκL =1 The last condition gives κL =2πn, n =0 ±1 ±2. This immediately gives a formula for the energy very similar to that for the box with ends, but with the replacement n →2n: E n = (2n) 2 h 2 8mL 2  (4.9) wherethistimen =0 ±1 ±2 The corresponding wave functions are ψ 0 =  1 L for n =0 ψ n>0 = Asin 2πn L x +B cos 2πn L x ψ n<0 =−A sin 2π|n| L x +B cos 2π|n| L x 12 And neglecting such effects as the particular shape of the benzene (curvature, etc.). 13 This is a set of homogeneous linear equations. 152 4. Exact Solutions – Our Beacons Since ψ n>0 and ψ n<0 correspond to the same energy, any combination of them also represents an eigenfunction of the Schrödinger equation corresponding to the same energy (Appendix B on p. 895). Taking therefore as the new wave functions (for n = 0) the normalized sum and difference of the above wave functions, we finally obtain the solutions to the Schrödinger equation  0 ≡ψ 0 =  1 L for n =0  n>0 =  2 L sin 2πn L x for n>0  n<0 =  2 L cos 2πn L x for n<0 4.2.3 COMPARISON OF TWO BOXES: HEXATRIENE AND BENZENE Let us take an example of two molecules: hexatriene and benzene (i.e. the cyclo- hexatriene). Let us assume for simplicity that the length of the hexatriene L is equal to the perimeter of the benzene. 14 Both molecules have 6 π electrons (any of them). The electrons doubly occupy (the Pauli exclusion principle) three one- electron wave functions corresponding to the lowest energies. Let us compute the sum of the electron energies 15 (in the units h 2 8mL 2 , to have the formulae as compact as possible): • HEXATRIENE: E heks =2 ×1 +2 ×2 2 +2 ×3 2 =28, • BENZENE: E benz =2 ×0 +2 ×2 2 +2 ×2 2 =16. We conclude, that 6 π electrons in the benzene molecule correspond to lower energy (i.e. is more stable) than the 6 π electrons in the hexatriene molecule. Chemists find this experimentally: the benzene ring with its π electrons survives in many chemical reactions, whereas this rarely happens to the π-electron system of hexatriene. Our simple theory predicts the benzene molecule is more stable than the hexatriene molecule. And what about the electronic density in both cases? We obtain (Fig. 4.4.b,c) • HEXATRIENE: ρ(x) =2 × 2 L [sin 2 π L x +sin 2 2π L x +sin 2 3π L x], • BENZENE: ρ(x) =2 × 1 L +2 × 2 L [sin 2 2π L x +cos 2 2π L x]= 6 L . 14 This is to some extent an arbitrary assumption, which simplifies the final formulae nicely. In such cases we have to be careful that the conclusions are valid. 15 As will be shown in Chapter 8, this method represents an approximation. 4.3 Tunnelling effect 153 This is an extremely interesting result. The π-electron density is constant along the perimeter of the benzene mole- cule. No single and double bonds – all CC bonds are equivalent (Fig. 4.4.c). Chemical experience led chemists already long time ago to the conclusion that all the C–C bonds in benzene are equivalent. This is why they decided to write down the ben- zene formula in the form of a regular hexagon with a circle in the middle (i.e. not to give the single and double bonds). The FEMO method reflected that feature in a naive way. Don’t the π electrons see where the carbon nuclei are? Of course they do. We will meet some more exact methods in further chapters of this textbook, which give a more detailed picture. The π-electron density would be larger, closer to the nuclei, but all CC bonds would have the same density distribution, similar to the solution given by the primitive FEMO method. From (4.9) and the form of the wave functions it follows that this will happen not only for benzene, but also for all the systems with (4n +2)-electrons, n =1 2, because of a very simple (and, therefore, very beautiful) reason that sin 2 x +cos 2 x =1. The addition or subtraction of an electron makes the distribution non-uniform (Fig. 4.4.d,e). Also in six π electron hexatriene molecule, uniform electron density is out of the question (Fig. 4.4.b). Note that the maxima of the density coincide with the double bonds chemists like to write down. However, even in this molecule, there is still a certain equalization of bonds, since the π electrons are also where the chemists write a single bond (although the π electron density is smaller over there 16 ). Again important information has been obtained at almost no cost. 4.3 TUNNELLING EFFECT Is it possible to pass through a barrier with less energy than the barrier height? Yes. 4.3.1 A SINGLE BARRIER Let us imagine a rectangular potential energy barrier (Fig. 4.1.b) for the motion of a particle of mass m: V(x)= V for 0  x  a,withV(x)= 0 for other values of x (V is the barrier height). Let us assume that the particles go from left to right and that their energy E is smaller than V . This assumption will make it possible to study the most interesting phenomenon – tunnelling through the barrier. 17 In order to stress that 0  E  V letuswrite: E =V sin 2 β (4.10) 16 Where, in the classical picture, no π electron should be. 17 Another interesting question would be what will happen if E>V. This question will be postponed for a moment. 154 4. Exact Solutions – Our Beacons The x axis will be divided in three parts: region 1 −∞<x<0, region 2 0  x  a, region 3 a<x<∞. In each of these regions the Schrödinger equation will be solved, then the solu- tions will be stitched together in such a way as to make it smooth at any boundary. The general solution for each region has the form 18 (x) =Ae iκx +Be −iκx ,where A and B are the de Broglie wave amplitudes for motion to the right and to the left. The κ constant comes from the Schrödinger equation ∂ 2  ∂x 2 +κ 2  = 0, where κ 2 = 2mE ¯ h 2 for regions 1 and 3 and κ 2 = 2m(E−V) ¯ h 2 for region 2. Therefore, the wave functions for each region is:  1 (x) =A 1 e ix √ 2mE ¯ h +B 1 e − ix √ 2mE ¯ h  (4.11)  2 (x) =A 2 e −x cotβ √ 2mE ¯ h +B 2 e x cotβ √ 2mE ¯ h  (4.12)  3 (x) =A 3 e ix √ 2mE ¯ h +B 3 e − ix √ 2mE ¯ h  (4.13) The second equation needs a little derivation, but using eq. (4.10) this is straight- forward. In regions 1 and 2 we may have the particle going right or left (reflection), hence in these regions A and B are non-zero. However, in region 3 we are sure that B 3 =0, because there will be no returning particle (since no reflection is possible in region 3). Now, the coefficients A and B are to be determined (with accuracy up to a multiplicative constant) in such a way as to ensure that the wave function sections match smoothly. This will be achieved by matching the function values and the first derivatives at each of the two boundaries. 19 As the wave function has to be continuous for x =0andx = a we obtain the following equations A 1 +B 1 =A 2 +B 2  A 2 exp  − a cotβ √ 2mE ¯ h  +B 2 exp  + a cotβ √ 2mE ¯ h  =A 3 exp  ia √ 2mE ¯ h   18 This is the free particle wave function. The particle has the possibility (and, therefore, certain prob- ability) of going left or right. 19 The second derivative is discontinuous, because of the form of the potential function V(x) intro- duced. 4.3 Tunnelling effect 155 The continuity of the first derivative at x =0andx =a gives: i(A 1 −B 1 ) =cot(B 2 −A 2 ) cotβ  −A 2 exp  − a cotβ √ 2mE ¯ h  +B 2 exp  + a cot β √ 2mE ¯ h  =iA 3 exp  ia √ 2mE ¯ h   After introducing the abbreviations: k =exp  a cotβ √ 2mE ¯ h  and N =(1 −k 2 ) cos2β +i(1 +k 2 ) sin2β we obtain the following ratios of all the coefficients and coefficient A 1 : B 1 A 1 = k 2 −1 N  A 2 A 1 = k 2 (1 −exp(−2iβ)) N  B 2 A 1 = (exp(2iβ) −1) N  A 3 A 1 = 2ik sin2β exp(− ia √ 2mE ¯ h ) N  A current in region 3 towards the positive direction of the x axis may be defined as the probability density A ∗ 3 A 3 of the particle going right in region 3 times the velocity  2E m =  2 mv 2 2 m =v in this region. Therefore, the current passing through region 3 is equal to current A ∗ 3 A 3  2E m = 4k 2 A ∗ 1 A 1 sin 2 2β |N| 2  2E m  (4.14) Therefore, the ratio of the current going right in 3 to the current going right in 1 is equal to: D single = |A 3 | 2 |A 1 | 2 (the barrier transmission coefficient, in our case equal transmission coefficient to the probability of passing the barrier): D single = 4k 2 sin 2 (2β) NN ∗ (4.15) . atom orbitals will easily recognize the resemblance of the above figures to some of them (cf. pp. 180–185), because of the rule mentioned above. the length of the molecule (averaging the potential energy. prob- ability) of going left or right. 19 The second derivative is discontinuous, because of the form of the potential function V(x) intro- duced. 4.3 Tunnelling effect 155 The continuity of the first. the x axis within a section of length L equal to 148 4. Exact Solutions – Our Beacons Fig. 4.3. Examples of the wave functions for a particle in a square box, the quantum numbers (n 1 n 2 ) correspond

Ngày đăng: 06/07/2014, 09:20

Từ khóa liên quan

Tài liệu cùng người dùng

Tài liệu liên quan