Sie sind auf Seite 1von 23

Beyond Lewis Structures The building-block formalism of organic chemical structure, more commonly called the "Lewis dot

model," is a powerful schema for organizing organic molecules and identifying similarities between them. Coupled with the generalized stability trends, we can begin to reason from the seen to the unseen to make predictions about never-before-seen chemical reactions. However, there are many observations that the Lewis dot model cannot explain. Its explanatory power doesn't even come close to its organizational strength. In fact, most stability trends find their origin not in the Lewis dot model, but in a deeper theory of molecular structure: molecular orbital theory. In this chapter, we'll develop a version of molecular orbital theory useful for the student and everyday practitioner of organic chemistry. Realize that we will not disprove the Lewis dot model in this chapter! Molecular orbital theory is meant to enhance, not replace the Lewis dot model. Its organizational power still remains, and we will still rely on the building-block formalism to make connections between analogous structures in later discussions. Before diving in to the details of molecular orbital theory, let's explore some of the shortcomings of the Lewis dot model. What can't it explain, and why is an additional theory necessary? Consider the process in Figure 1a, rotation about a carbon-carbon double bond. The Lewis dot model stipulates that the four electrons in the double bond are shared between the two carbon atoms, but it can't explain why rotation about the double bond doesn't occur until the compound is heated to extremely high temperatures. Molecular orbital theory reveals that carbon-carbon multiple bonds cannot rotate without breaking.

Organic Chemistry/Evans

Figure 1. Observations that the Lewis dot model does not explain adequately. (a) Rotation about double bonds does not occur at room temperature, even though rotation around unhindered single bonds is rapid at room temperature. (b) One-step substitution reactions result in an inversion of configuration (two atoms appear to switch places) at the electrophilic carbon. (c) Carbocations substituted with more carbon atoms are more stable than less substituted cations. The single-step SN2 reaction in Figure 1b has properties that are not addressed by the Lewis dot model, too. Displacement of bromide by hydroxide leads exclusively to an inversion of configuration at the central carbon atom. This observation suggests a particular trajectory for hydroxide as the substitution takes place, but the Lewis dot model offers no dominant path. The spatial aspects of the reaction are best explained by molecular orbital theory--in fact, simple steric considerations may argue against the observed configuration of the product! Finally, Figure 1c shows the relative stability of two carbocations. More substituted cations are more stable than those that are less substituted, other things being equal. Yet, aside from the somewhat handwavy explanation of steric hindrance due to flanking CH bonds, the Lewis dot model cannot account for the exceptional stability of the left-hand (tert-

Organic Chemistry/Evans

butyl) cation. Orbital interactions, which find their theoretical basis in molecular orbital theory, elegantly explain this ubiquitous observation. To account for all of the observations in Figure 1, we need a model that suggests the spatial positions and energies of electrons within a molecule. Molecular orbital theory provides this information--in fact, an orbital is just an electronic container with an associated energy value. In the next section, we'll introduce the fundamentals of the orbital concept. Throughout the following sections, keep in mind our goal of describing in detail the electronic structure and reactivity of organic molecules. *** Introduction to Orbitals In this section, we'll begin to extend the Lewis model by expanding our conception of the electron. While the Lewis dot model stipulates that electrons sit, isolated, on or between atoms, the quantum mechanical truth is far more interesting: electrons may be delocalized over large regions of space, and even across several atoms! Belonging to each electron in a molecule is a function called an orbital, which describes all of its properties according to the principles of quantum mechanics. Strictly speaking, an orbital is a function over all space that specifies the probability of finding an electron at a given point.1 Orbitals for atoms and molecules take into account the positions of nuclei and other electrons nearby. In practice, probability values rapidly approach zero at certain distances from the nucleus. For this reason, we can focus only on a "slice" of space, a container that encloses all of the points for which the probability of finding the electron is larger than some cutoff value.2 Organic chemists most often think of orbitals in this way, as containers in which electrons are likely to be found. A single orbital can contain up to two electrons of different electronic spin. We typically refer to the different spins of the electron as "spin up" and "spin down" or +1/2 and 1/2. To depict the occupancy, energy, and shape of orbitals, chemists use orbital diagrams. A simple orbital diagram is shown in Figure 2. Notice that the number of electrons, energy, and appearance of the orbital are all shown on the orbital diagram. A horizontal line is drawn to indicate the position of the orbital on a vertical energy scale. Half arrows up and down are drawn to show how many electrons are in the orbital, and what their spins are. Often, the shape of the orbital will be omitted or drawn in a separate picture of the molecule

Organic Chemistry/Evans

under study.

Figure 2. The three essential features of every orbital, with an example of an orbital diagram. Orbitals can take on either positive or negative values over space. This seems at odds with our earlier definition of the orbital as a distribution of probabilities: how can a probability be negative? We now need to expand our previous definition: the magnitude of the orbital is a probability, and its sign refers to a property called phase. Phase is important when orbitals combine or overlap, because they do so like waves, enhancing one another in some regions and canceling one another out in others. In pictures of orbital shapes, positive and negative phases are indicated either with different colors or with shading. In Figure 2, for example, we might choose the unshaded area to represent a region of positive phase and the shaded area to represent a region of negative phase. Our exact choice is arbitrary, but it's important to keep in mind that the shaded and unshaded regions are places where the signs of the orbital are different. We'll revisit phase shortly, when we discuss how orbitals combine with one another. Orbitals are special functions that apply to the electron. But where do the functions themselves come from? What's the origin of the shapes and energies of orbitals? We won't follow this line of inquiry too far, except to say that the orbital belonging to an electron depends on the electron's environment. Designations like "atomic" and "molecular orbitals" reflect the environment of the electron. Atomic orbitals belong to electrons in atoms,

Organic Chemistry/Evans

molecular orbitals belong to electrons in molecules, et cetera. The nature of the environment and a quantum mechanical relation called the Schrdinger equation are used to solve for the form of the orbital. We'll largely treat the shapes and energies of atomic orbitals as axiomatic--that is, we'll present them without proof. Such proofs are generally reserved for physical chemists...and organic chemists are glad to give them up! Watch The Orbital Concept Let's begin with a look at the atomic orbitals (AOs) that characterize the first- and second-row elements. There are five that are important for organic chemistry: the 1s, 2s, 2px, 2py, and 2pz atomic orbitals. Their shapes and relative energies are shown on an orbital diagram in Figure 3. At the center of each 2p orbital, where the nucleus sits, we can identify a node, where the phase of the orbital changes. At the node, the value of the orbital is zero. These five orbitals form a sort of scaffold, into which we can add electrons to characterize the organic atoms.

Figure 3. The atomic orbitals on an orbital energy diagram, with

Organic Chemistry/Evans

electronic configurations of the first- and second-row elements. We use the term electron configuration to refer to how an atom's electrons occupy the atomic orbitals. In the blue box in Figure 3.3, the configurations of the first- and second-row elements are provided. The first number in each orbital's name is its principal quantum number, a measure of its energy (as principal quantum number goes up, energy does too). The letters s and p indicate the orbital's subshell (an indicator of orbital shape and energy). Finally, the superscripted 1's and 2's show the number of electrons in each AO. Notice that electrons occupy the most stable orbitals first, followed by more unstable orbitals. The three 2p orbitals are all degenerate--that is, they have the same energy. Degenerate orbitals are filled one electron at a time, so that two electrons are not paired up until they must be. Figure 4 illustrates how electrons fill the AOs of boron, carbon, nitrogen, and oxygen. Notice that two electrons don't occupy the same orbital until it's unavoidable. This method for filling orbitals with electrons, which holds regardless of the type of orbital (atomic, molecular, etc.) is called Hund's rule.

Figure 4. Orbital energy diagrams for B, C, N, and O atoms. Hund's rule states that electrons should not be paired up until no other placement is possible.

Organic Chemistry/Evans

Let's take a breather--what's the point of all this atomic orbital mumbo-jumbo? With some notable exceptions, we won't deal directly with the atomic orbitals once we've fully developed molecular orbital theory. However, atomic orbitals are quite literally the building blocks of molecular orbitals, which are our ultimate endgame. We'd like to understand the electronic structures of molecules, but to do so, we need to recognize how electrons are arranged in atoms. In particular, the molecular orbitals we tend to care about are built from the valence atomic orbitals, an atom's occupied orbitals with highest principal quantum number.3 Early in this section, we noted that the shape and energy of an electron's orbital depends on the system of which it's a part. Atomic orbitals pertain to atomic systems, molecular orbitals to molecular systems, etc. An orbital is a solution to a problem that takes the system into account. Thus, from first principles, we might imagine that molecular orbitals, the solutions to molecular problems, are unrelated to atomic orbitals, the solutions to atomic problems. Yet, remarkably, we find in practice that linear combinations (weighted sums) of the atomic orbitals approximate solutions to the problem of describing electrons in molecules. This approximation is the cornerstone of the linear combinations of atomic orbitals-molecular orbitals (LCAO-MO) method, which we'll explore in the context of dihydrogen in the next section. Watch The Atomic Orbitals *** Molecular Orbitals of Dihydrogen Our goal so far has been to describe the spatial positions and energies of electrons in atoms and molecules, using the principles of quantum mechanics as a foundation. Quantum mechanics sets up the problem of determining an electron's orbital as a kind of physical optimization problem: what spatial arrangement of electrons leads to the most stable system? In the last section, we were introduced to the atomic orbitals, which are solutions for atomic systems of the first and second rows of the periodic table. Thankfully, we do not need to start over at square one to determine the orbitals of electrons in molecules! Molecular orbitals are well approximated by linear combinations of the atomic orbitals associated with the molecule's atoms.

Organic Chemistry/Evans

Keep our goal in mind throughout this section. We want to answer the following questions: Where are the electrons in a molecule located in space? What are their energies? We've already seen the containers that hold electrons in atoms--the 1s, 2s, and 2p atomic orbitals. Molecular orbitals are constructed from these building blocks, and have spatial and energetic properties that reflect their construction from atomic orbitals. In fact, there are multiple ways to carry out the process of building molecular orbitals, but we will adopt a localized molecular orbital theory approach using the ideas of hybridization and localized MOs. Localized MO theory breaks from the "canonical" mold used by physical chemists, but will serve us well throughout future discussions. Let's begin with a very simple case: the molecular orbitals of the dihydrogen molecule, H2. The hydrogen atom possesses a single electron in a 1s atomic orbital. When two hydrogen atoms come together, a molecule containing a total of two electrons results (H2). Thus, we might imagine that only one molecular orbital is needed to accommodate these two electrons. However, this conclusion ignores the important fact that hydrogen's valence AO (1s) is unfilled. Both the hydrogen atom and the dihydrogen molecule can accept electrons. We could imagine giving one more electron to the hydrogen atom to create the hydride anion, H. Then, we could combine two H atoms together to create the fantastical dianion, H22. Loading up the valence atomic orbital of H with electrons shows us that we need two molecular orbitals for the neutral H2 molecule, even though the molecule itself only possesses two electrons. H2 can accept two additional electrons, and we need a place to put those! In general, remembering that we need a place to put any electrons that may enter the molecule, we can conclude the following: The total number of atomic orbitals possessed by all the atoms of a molecule equals the molecule's total number of molecular orbitals. Figure 5 below shows the molecular orbitals of dihydrogen at the

Organic Chemistry/Evans

center of an orbital energy diagram. The shapes of the molecular orbitals are determined by quantum mechanical principles that we won't concern ourselves with here. What we should notice is the relationship between orbital shape and energy--notice that the MO with lobes of two different phases (and a node in between) is higher in energy than the orbital lacking any sign changes. Mathematically, the higher-energy orbital is the result of subtracting the two 1s AOs, and the lower-energy orbital is the result of adding the two AOs. When we add two 1s orbitals of the same phase, they reinforce one another in the region between the nuclei. When we add orbitals of opposite phase, however (i.e., subtract two AOs of the same phase), they tend to cancel one another out between the nuclei, and at some point, a node results. Figure 5 is probably the simplest example of the linear combinations of atomic orbitals-molecular orbitals (LCAO-MO) method. The dotted lines from the atomic to the molecular orbitals indicate contributions of the AOs to each MO.

Figure 5. A molecular orbital energy diagram for H2, with isolated H atoms (and atomic orbitals) on the periphery and the molecule (and its MOs) in the center. Because the higher-energy orbital lacks electron density between the

Organic Chemistry/Evans

10

nuclei, we call that orbital antibonding. Antibonding orbitals have energies that are higher than those of the isolated atoms. Bonding orbitals, on the other hand, have enhanced electron density between the nuclei and lower energies than the separated atoms. Bonding MOs involve constructive overlap of AOs, overlap of two lobes of the same phase. Antibonding orbitals result from destructive overlap--when lobes of opposite phase coincide in space. The simple example of dihydrogen from this section has introduced us to the idea that molecular orbitals are built as weighted sums and differences of atomic orbitals. In future discussions, we'll create molecular orbital diagrams like Figure 5 using other types of atomic orbital building blocks that sit on nearby atoms. Let's end this section with a quandary. We've seen that molecular orbitals may be built from the 1s, 2s, and 2p atomic orbitals. We've also seen that the 2p orbitals are at right angles to one another. The angles of the directional 2p orbitals seem inconsistent with the bond angles we saw in the section on molecular geometry. How can we reconcile these two ideas, which seem to be at odds with one another? We'll see how the idea of hybridization solves this problem in the next section. Watch Linear Combinations of Atomic Orbitals *** Atomic & Molecular Orbitals in Organic Molecules Applying the LCAO-MO method of the last section to large organic molecules produces complex, delocalized molecular orbitals. Such MOs tell us little about how we should expect molecules to behave. Furthermore, since delocalized MOs bear little resemblance to the lines and dots of Lewis structures, it can be difficult to make connections between the two. To get around this problem, we'd like an orbital theory that yields localized MOs that look and "feel" like the elements of Lewis structures we already know and love. In this section, we'll learn the six localized molecular orbitals that correspond directly to bonds and lone electron pairs in Lewis structures. Our goal is to describe the shapes and energies of electron sources and sinks in more detail, to paint a more vivid picture of electronic structure in organic molecules. At the end of the last section we confronted a quandary: is it possible

Organic Chemistry/Evans

11

to reconcile the geometry of the 2p atomic orbitals (90 bond angles) with the observed geometries of organic compounds (tetrahedral, trigonal planar, and linear)? The answer is a resounding "yes," thanks to Linus Pauling's concept of hybridization. To account for the observed geometries of organic compounds, Pauling proposed that before combining with orbitals on other atoms, atomic orbitals can "hybridize" to produce a new set of atomic orbitals for bonding. The process of "hybridization" is essentially a kind of on-atom linear combination. We can take bits and pieces from the different atomic orbitals to construct the hybrid atomic orbitals. Don't worry about how exactly this is done--it's important just to recognize the final result of hybridization, the three sets of hybrid AOs in Figure 6.

Figure 6. The three sets of hybrid orbitals and their corresponding generalized building blocks. The relative energies of the hybrid atomic orbitals are straightforward to understand.4 In the last section, we saw that the 2p atomic orbitals are higher in energy than the 2s orbital. Thus, we might expect hybrid orbitals made of a greater percentage of 2p AOs to be higher in energy than those

Organic Chemistry/Evans

12

with less 2p character. Experiments support this idea--the stability of electrons held in hybrid AOs follows this trend: (highest energy) 2p > sp3 > sp2 > sp > 2s (lowest energy) Stated another way, hybrid AOs of greater s character are lower in energy than those with less s character. Does this lower energy automatically imply greater stability? No! The lower-energy hybrids are stable when filled, but unstable without electrons. For this reason, cations and other unsaturated species that possess empty hybrid orbitals are usually unstable. There is a profound link between the nature of an atom's building block within a molecule and its hybridization. From Figure 6, notice that hybridization depends on its geometry: tetrahedral atoms are sp3hybridized, trigonal planar atoms are sp2-hybridized, and linear atoms are sp-hybridized.5 Furthermore, the number of ! bonds on an atom corresponds to the number of AOs used to form its hybrids: 4 ! bonds = 1 s + 3 p = sp3, et cetera. Multiple bonds, as we will soon see, can be understood as arising from interactions between unhybridized, leftover 2p orbitals. For instance, sp2 hybridization leaves behind one 2p orbital for multiple (") bonding. sp Hybridization leaves behind two 2p orbitals for " bonding. Examining the building blocks corresponding to these hybridization states, we see one and two multiple bonds, respectively. That's not a coincidence! The sp2 and sp hybrid orbital sets are shown in Figure 7 with their leftover 2p orbitals.

Organic Chemistry/Evans

13

Figure 7. Hybridization may leave behind unused 2p orbitals, if less than four ! bonds are needed. Leftover 2p orbitals may be used for " (multiple) bonding or to hold nonbonding electrons. The hybrid atomic orbital sets will serve as our starting point for thinking about the molecular orbitals of large organic compounds. Localized molecular orbitals are built either from the hybrid AOs themselves, in isolation (we call such molecular orbitals non-bonding), or from bonding and antibonding combinations of the hybrids. At this stage, it's important to realize that we can treat each ! (single) bond in an organic compound like we did the hydrogen atom, using hybrid orbitals in place of 1s orbitals. For each ! bond, there is a ! bonding orbital, which holds the electrons of the bond, and a !* antibonding orbital, which reflects the ability of the bond to break upon the addition of two more electrons. The bonding MO involves constructive overlap of the hybrids, and the antibonding MO destructive overlap. Figure 8 shows a simple orbital energy diagram for the carboncarbon ! bond in ethane (C2H6).

Organic Chemistry/Evans

14

Figure 8. A simple orbital energy diagram for the CC bond of ethane. The bond is associated with two molecular orbitals: a bonding ! orbital of low energy, and an antibonding !* orbital of high energy. The ! molecular orbital shows us where the electrons of the bond are likely to be found in space. The energy of this orbital reflects the reactivity of the electrons in the bond as an electron source--a clean, intuitive idea. But how should we interpret the empty antibonding orbital? What does an antibonding orbital "mean" from the molecule's perspective? Put most concretely: how does the nature of the !* orbital affect the bond's behavior (structure and reactivity)? Spatially, the !* orbital shows us where incoming electrons from a source are likely to go. Incoming electrons will tend to approach the large lobes of the antibonding orbital. Like unfilled hybrid AOs, antibonding MOs are most stable when high in energy. Summing up, the nature of the !* orbital reflects the bond's potential as an electron sink! Since reactions require both an electron source and sink, we must keep the importance of antibonding orbitals in mind when studying organic reactivity--the sink is an essential piece of the puzzle. ! Bonding is advantageous for molecules because, as we can see from Figure 8, the overall energy of electrons is lowered in the process. But

Organic Chemistry/Evans

15

! bonding also seems to present certain problems--for instance, how can hybrids overlap to form multiple bonds? Using more than one hybrid orbital to describe a multiple bond seems out of the question. Still, the 2p orbitals left behind on sp2- and sp-hybridized atoms don't look appropriately positioned for orbital overlap. What gives? Evidently, the parallel arrangement of 2p orbitals is good enough to establish a bond between adjacent atoms. Figure 9 shows the idea for the simplest hydrocarbon that contains a double bond, ethylene.

Figure 9. Multiple bonds are the result of "side on," parallel, or "-type overlap of adjacent 2p orbitals. Just as in the ! bonding case, a bonding " orbital and an antibonding "* orbital result from constructive and destructive overlap. Our spatial and energetic interpretations of the " and "* orbitals are identical to those of the ! and !* orbitals. To reiterate, the " orbital reflects the potential of the double bond to serve as an electron source, and the "* orbital reflects the potential of the double bond to serve as an electron sink. Importantly however, the side-on overlap characteristic of " orbitals is weaker than the head-on overlap involved in ! bonding. As a result, " and

Organic Chemistry/Evans

16

"* MOs tend to be closer in energy to their atomic orbital building blocks than ! orbitals. Consequently, two important trends occur: " Bonding MOs are higher in energy than ! MOs. "* Antibonding MOs are lower in energy than !* MOs. Taken together, what do these trends suggest about the relative reactivity of single and multiple bonds? In general, multiple bonds are more reactive (both as sources and sinks) than single bonds. Electrons in " bonds are higher in energy than electrons in ! bonds; unfilled "* orbitals are lower in energy than !* orbitals. In addition, we can see from Figures 8 and 9 that the spatial positions of " and ! electrons differ. ! Electron density can be found along an axis connecting the atoms, while " electron density is found above and below such an axis. In fact, " MOs possess a node coinciding with the axis connecting their atoms. Finally, we need to address nonbonding electrons and orbitals, which are not involved in interactions with orbitals on adjacent atoms. In the absence of resonance interactions, lone electron pairs can be found in isolate hybrid AOs. We will refer to these as n molecular orbitals (for "nonbonding"). Figure 10 depicts three examples of n orbitals in common compounds. Notice that each n orbital is characterized by a particular hybridization, since the n MO is really just a hybrid atomic orbital. The n orbital reflects the potential of the atom to serve as an electron source via a lone pair sitting on the atom.

Figure 10. Nonbonding n orbitals in common organic molecules. Most commonly, n orbitals are just hybrid atomic orbitals. Building blocks bearing fewer than 8 total electrons must possess an empty nonbonding MO, which we call an a orbital (for "atomic"). What are

Organic Chemistry/Evans

17

the atomic orbital constituents of a MOs? Empty hybrid AOs present an energetic problem: they're low in energy and thus tend to be unstable when lacking electrons. To get around this problem, molecules without geometric constraints adopt geometries that allow them to leave high-energy 2p orbitals unfilled. Thus, in the vast majority of cases, a orbitals are just empty atomic 2p orbitals. When you spot a building block with 6 or fewer total electrons, take note that an empty a orbital is present on the atom! This MO reflects the potential of the atom itself to serve as an electron sink. Empty a orbitals are typically the lowest energy (and most unstable) unfilled orbitals one finds in organic compounds. This fact is unsurprising when we consider that building blocks bearing a orbitals lack an octet of electrons. Figure 11 provides two examples of empty a orbitals: the typical a = 2p case, and a case in which the empty orbital must be a hybrid AO (based on the geometry of the building block, which demands sp2 hybridization).

Figure 11. Empty atomic, a orbitals in organic molecules. a MOs are most commonly just 2p atomic orbitals; however, geometric constraints may force a hybrid AO to be empty. Notice that both cationic building blocks have 6 total electrons. With the a orbital, we've reached the sixth and last of the localized MO classes. Figure 12 provides a summary of the orbital shapes of the six classes and the structural elements to which they correspond. These shapes both confirm our intuition and suggest some intriguing new ideas. On the confirmatory side, notice that electron sources tend to be concentrated between nuclei, where the Lewis model suggests we should find electrons. More interestingly, the !* electron sink is primarily located outside of the space between the nuclei. Incoming electrons will most likely approach the outskirts of a bond, not between the atoms. Trigonal planar atoms lacking an octet of electrons (thus bearing an a MO) will be

Organic Chemistry/Evans

18

approached by electrons perpendicular to the molecular plane. Consider the " MOs--notice that they depend on a parallel alignment of 2p orbitals. Rotation away from the parallel alignment ruins "-type overlap, so double bonds must remain planar. All kinds of interesting spatial ideas come to light!

Figure 12. Shape and occupancy of the six classes of localized molecular orbitals. The filled orbitals reflect the potential of the structure as an electron source; the empty orbitals reflect the structure's potential as an electron sink. The relative energetics of these six classes are also extremely important to keep in mind. These form a kind of stability trend that allows us to quickly pinpoint the most reactive sources and sinks within a molecule. We touched on this trend in Figure 13, but it's worth bringing up again here, now that we've seen the shapes of the localized MOs. Figure 14 depicts the most common relative energies of the localized MOs. Based on Figure 14, we should expect n orbitals to be the most reactive sources and a orbitals to be the most reactive sinks. By identifying the localized MOs present in a molecule and using Figure 14 to predict their relative energies, we can predict how molecules will behave in a very powerful, general way.

Organic Chemistry/Evans

19

Figure 14. Typical relative energies of the localized molecular orbitals. To drive the point home once more, low-energy unfilled orbitals and high-energy filled orbitals are reactive. Watch Localized Molecular Orbital Theory *** Effects of Electronegativity & Charge Figures 12 and 14 lay out the shapes and relative energies of the six classes of localized MOs. But a simple problem reveals that those figures don't tell us the whole story. Consider Figure 15 below, which depicts a good nucleophile (thiolate anion) in the presence of a compound containing C=O and C=N bonds. Based on ideas from the last section, the "* orbitals of the C=O and C=N bonds should be the electrophile's best electron sinks. Can we predict which of the two sinks is better?

Organic Chemistry/Evans

20

Figure 15. Two possible courses of action: thiolate can add either to the C=O or C=N bonds. Can we used localized MO theory to predict which of the two "* orbitals is more reactive? We can, if we first recognize that the identity of the atoms that make up the two double bonds are different. Where we see a nitrogen in one double bond, an oxygen sits in the other. We need to understand how atom type affects the shapes and energies of molecular orbitals. Even more specifically, we need to understand the relationship between electronegativity and the properties of orbitals. How does electronegativity affect the energies of the atomic orbitals? How are orbital shapes influenced by electronegativity? As we move from left to right across the periodic table, orbital energies decrease. Rather intuitively, more electronegative atoms are associated with lower energy orbitals. Another way of saying this is that more electronegative atoms are more electrophilic, or that they tend to be associated with more electrophilic orbitals. Put yet another way, more electronegative atoms make worse nucleophiles (electron sources). Whereas carbanions are extremely nucleophilic (electron-donating) molecules, fluoride anion is hardly nucleophilic at all. Applying this idea to the problem in Figure 15, we can predict that the "* orbital associated with the C=O bond ("*CO) is lower in energy than the "* orbital associated with the C=N bond ("*CO). Since this orbital is unfilled, the C=O bond ought to be more reactive than the C=N bond.6 Molecular orbitals may involve identical atom types that differ only in their charge. Thus, to finally complete our understanding of the effects of

Organic Chemistry/Evans

21

molecular structure on orbital energy and shape, we need to learn how charge influences orbitals. Think of charge as a specialized case of electronegativity, making use of the classic maxim that "opposite charges attract and like charges repel." Positive charge on an atom boosts its electronegativity over its neutral counterpart--positively charged atoms hold electrons more tightly than neutral atoms (opposites attract). Negative charge has the opposite effect, and lowers the electronegativity of the atom relative to its neutral counterpart (like charges repel). Consider the example in Figure 16.

Figure 16. Which of the two C=N double bonds ought to be more reactive under these conditions? Since positive charge increases electronegativity, we can think about the positively charged nitrogen atom in the same way we thought about oxygen in the first example from this section. It's more electronegative than neutral nitrogen, so the energy of its "* orbital is lower. Consequently, the C=N+ double bond ought to be more reactive than the C=N double bond. Organic chemists confirm predictions like these on a daily basis with experiments! Electronegativity's effects on orbital shape are intuitive to understand as well. Polarized MOs are built using atomic orbitals from atoms of very different electronegativity, and consist of lobes of different sizes on each atom. As we might expect, filled orbitals tend to have large lobes on electronegative atoms, since electrons tend to spend their time around atoms that are hungry for them. Unfilled orbitals, on the other hand, tend to have larger lobes on less electronegative atoms. We will revisit this idea

Organic Chemistry/Evans

22

when we discuss frontier molecular orbital theory and the fundamentals of reactivity. For now, it's most important for us to understand how electronegativity affects orbital energies. Our intuitive ideas about electronegativity map nicely onto orbital energy ideas: electronegative atoms possess low orbital energies and thus stabilize electrons. Treating charge as a kind of electronegativity booster or restrictor, we can treat charged atoms using this same idea. Armed with this principle, we can easily compare the electronic viability of two possible reaction pathways, even if the pathways in question differ only in the charge or element type of the atoms involved. 1. Orbitals are interchangeably called "wavefunctions." We'll use the term "orbital" whenever possible, but you should be aware that these two terms are often used interchangeably. 2. Because we think of orbital shapes as containers for electrons, we commonly say that an electron is "in" an orbital. 3. For example, in Figure 3.4 the valence AOs are the 2s level and all three 2p levels. 4. See the subsection addressing hybridization in the "Generalized Stability Trends" section. 5. This relationship between geometry and hybridization always holds true. However, it's important to remember that "geometry dictates hybridization, not the other way around." (Scott Denmark) Hybridization follows from geometry, and not necessarily from the apparent bonding network of the atom. Benzyne is a classic example of this idea: the triple bond of benzyne suggests sp hybridization for the two atoms involved; however, geometry demands that they be at least approximately sp2-hybridized. The triple bond of benzyne more closely resembles an sp2-hybridized diradical than a "classical" triple bond. 6. It's important to recognize this thought process as highly context dependent. Say we treated the organic compound in Figure 15 with an acid (HA) instead. Acids are electrophilic species, so we're interested in the organic compound as a nucleophile. Comparing the n orbitals on oxygen

Organic Chemistry/Evans

23

and nitrogen, we can conclude that the n orbital on less electronegative nitrogen is higher in energy than the n orbital associated with oxygen. Thus, nitrogen should be the better nucleophile (or base) in this context. For every reaction you encounter, ask yourself which of the starting materials are nucleophiles (electron sources) and electrophiles (electron sinks). As these examples show, mixing up these concepts can have damaging effects on predictive ability.

Das könnte Ihnen auch gefallen