Sciencemadness Discussion Board

Quantum Mechanics/Wave Mechanics in Chemistry Beginners Thread

 Pages:  1  ..  3    5    7  ..  9

Darkstar - 4-11-2015 at 11:58

Quote: Originally posted by aga  
Are the OC answers open to everyone, and should be U2U'd to you Darkstar, same as for the QM answers to blogfast ?


Those questions were primarily intended for you personally. You can just post your answers here in the thread if you want. I'll post the correct answers afterwards. This way, anyone who's been quietly following this thread can formulate their own answers and then check to see if they were right or wrong after I post the correct ones.

aga - 4-11-2015 at 13:06

I'm honoured !

OK. Here goes.

OC Q1

Step 1

The OH- has one electron too many.

The ethyl acetate seems to have all of it's atoms quite happy, with no free electrons.

The Carbon at the second position from the left will be having the electron density pulled away from it towards the more electronegative Oxygen (orbital probability density closer to the O than the C) effectively making it relatively electron-poor.

That should make the O of the OH- the nucleophile and the 2nd C of the ethyl acetate the electrophile.

Step 3

The O on the ethanol has the extra electron so must be the nucleophile.

The right-hand H on the sodium acetate (where's the Na ?) couldn't give a shit, as it has it's orbital full and happy, so the electrophile must be where the electron ends up, which is on the Right-most O of the sodium acetate.

OC Q2

Hmm. Will give it a go.

Step 1

The OH- bonds with the 2nd C giving it a three-O problem, causing the pi bond to the double bonded O to be disrupted.

This leaves the orbitals all in a mess, and an electron homeless.

Step 2

The homeless electron ends up with the right-hand O for no good reason i can think of, other than that it has a spare p orbital for it to half-occupy.

Step 3.

Being less electronegative due to this extra electron, the O on the ethanol structure is ejected, along with the rest of the ethanol structure.

The spare electron is sucked up by the right-most O on the ethyl acetate (via the adjacent H) for no good reason that i can think of.

Extra Credit at the B&D Education Store
Erm, dunno.

After Step #1 there's already an electron excess, so further excess electrons can't have much effect ?

Needs some H+ hanging about to knock out that extra electron ?

aga - 4-11-2015 at 14:44

Might be an idea to go U2U on this with a time limit, same as the QM Q/A thing.

The thread might fill up with junk otherwise.

Appologies if what i recently posted also constitutes Junk.

Darkstar - 4-11-2015 at 20:13

Quote: Originally posted by aga  
Might be an idea to go U2U on this with a time limit, same as the QM Q/A thing.


I'm just going to post the answers now so we can move on. The real purpose of this quiz was to gauge where you are at the moment so I can determine if there are any concepts that we still need to work on before we get a bit more advanced. I'm going to go ahead and post the correct answers below and then personally address your answers in a separate post. The correct answers are:

Quote:
Answers to OC Quiz #1

Question 1 - Identify the electrophile and nulcleophile in steps one and three of the mechanism.

Correct Answer: In step one, the nucleophile is the hydroxide ion and the electrophile is the carbon of the ester group. In step three, the nucleophile is the ethoxide ion and the electrophile is the hydrogen of the carboxylic acid group.

Question 2 - Briefly explain what is going on in each step of the mechanism.

Correct Answer: In the first step, the nucleophilic hydroxide ion attacks the electrophilic carbon of the ester group by donating one of its lone pairs to the electron-deficient carbon center. This breaks the pi-bond between carbon and the carbonyl oxygen, creating the negatively-charged tetrahedral intermediate seen in step two.

In the second step, a lone pair comes down from the top oxygen to form a new pi-bond with the electron-deficient carbon, simultaneously breaking the sigma bond between carbon and the ethoxy group and allowing the latter to leave as an ethoxide ion.

In the last step, the ethoxide ion grabs the acidic proton off of acetic acid, forming an acetate ion and a neutral ethanol molecule.

Extra Credit Question - Explain why the reaction is virtually irreversible under basic conditions.

Correct Answer: The second step of the mechanism is reversible because it is possible for the pi-bond formation to end up breaking the bond between carbon and the hydroxyl group instead, allowing it to leave as a hydroxide ion. This gives back the original ethyl acetate molecule and the reaction starts all over again. Similarly, step three is also reversible because there are two possible electrophiles for the nucleophilic ethoxide ion to attack: the hydrogen and carbon of the carboxylic acid group, both of which are electron-deficient.

If ethoxide attacks the carbon, the pi-bond breaks once again to return us back to the charged tetrahedral intermediate in step two. On the other hand, if ethoxide attacks the proton, a much weaker nucleophile is formed--ethanol. And because there's now a negative charge on the acetate ion, it is no longer electrophilic enough to be attacked by most nucleophiles--especially not one as weak as ethanol--and is protected by the electrostatic repulsion caused by the negative charge. Thus as long as the conditions are moderately basic, the acetate ion will remain deprotonated and unable to undergo further nucleophilic attack.


[Edited on 11-5-2015 by Darkstar]

Darkstar - 4-11-2015 at 20:45

Now that I've given out the official answers, I would like to make a few comments regarding your answers:

Quote: Originally posted by aga  
The Carbon at the second position from the left will be having the electron density pulled away from it towards the more electronegative Oxygen (orbital probability density closer to the O than the C) effectively making it relatively electron-poor.

That should make the O of the OH- the nucleophile and the 2nd C of the ethyl acetate the electrophile.


Correct!

Quote: Originally posted by aga  
The O on the ethanol has the extra electron so must be the nucleophile.The right-hand H on the sodium acetate (where's the Na ?) couldn't give a shit, as it has it's orbital full and happy, so the electrophile must be where the electron ends up, which is on the Right-most O of the sodium acetate.


The first part is correct about ethoxide being the nucleophile; however, the electrophile in the third step is the hydrogen! Just because it has a full orbital doesn't necessarily mean that it's happy. Like the carbon in the first step, hydrogen also has a fairly strong partial positive charge on it due to oxygen's electronegativity, making it a prime target for nucleophiles. Not to mention that the acetate ion's negative charge isn't actually localized on oxygen like it appears above. Because of resonance, the negative charge on oxygen that results from removing that proton is relatively weak. This makes it easier to remove the hydrogen because it is less strongly attracted back to the oxygen.

I'll do a brief lesson soon on what makes certain hydrogens more "acidic" (easily removable) than others, but for now just think of it as having to do with the degree that the electron density is being pulled away from them. The closer the electron density is to the hydrogen, the stronger it will be attracted to it and the harder it will be to pull away. On the other hand, when the electrons are being pulled far away from hydrogen, it becomes easier to remove because its attraction to those electrons (and their attraction back) is weaker.

In the case of carboxylic acids (like acetic acid), the negative charge that ends up on oxygen when the proton leaves is rather weak due to electron delocalization. Because of this, the hydrogen is quite easily removed because it is less attracted back as it leaves. Which is also why ethoxide is a MUCH stronger base than acetate, as the negative charge on ethoxide is more or less completely localized on oxygen. And because it's attached to an electron-donating carbon chain, ethoxide's negative charge ends up being extremely strong, even stronger than hydroxide's. This is what's known as a superbase. These kinds of bases can't exist in water because they instantly grab a proton off of water to create a weaker hydroxide ion that isn't strong enough to grab its proton back.

Also, as far as sodium and acetate ions are concerned, there aren't any shown in the mechanism. In the last step, acetic acid is deprotonated to give the acetate ion, which is only shown as one of the products in the overall reaction above the mechanism. I didn't bother to show the sodium ions in the mechanism, but they are always around, floating in solution. If the solution were evaporated, the oppositely-charged sodium and acetate ions would remain and counter one another, giving an ionic salt (sodium acetate).

Quote: Originally posted by aga  
The homeless electron ends up with the right-hand O for no good reason i can think of, other than that it has a spare p orbital for it to half-occupy.


As I mentioned in the answer section, it can end up with either oxygen (hydroxide or ethoxide). The only difference is that, when it ends up with ethoxide, in the event that it grabs a proton to give an acetate ion and an ethanol molecule, the reaction is over.

Quote: Originally posted by aga  
After Step #1 there's already an electron excess, so further excess electrons can't have much effect ?


Remember that there's really a partial negative charge (-1/3) on the carbon of the acetate ion due to electron delocalization. Electrostatic repulsion would prevent any further nucleophilic attacks on carbon.

[Edited on 11-5-2015 by Darkstar]

aga - 5-11-2015 at 01:00

Oh !

So the whole resonance thing is imagining (even calculating!) where the electron densities are on the molecule, especially the active bits, and the relative electrostatic charges that implies ?



blogfast25 - 5-11-2015 at 19:53

Quote: Originally posted by aga  
Oh !

So the whole resonance thing is imagining (even calculating!) where the electron densities are on the molecule, especially the active bits, and the relative electrostatic charges that implies ?




Basically, when it comes to reaction mechanisms, the whole point of QChemistry is working out the electron probability densities of molecules. These tell us where electrophilic (Lewis acids) or nucleophilic (Lewis bases) transformative attacks can take place: REACTIONS!

PS: don't forget Q 6!

aga - 6-11-2015 at 15:06

Cool !

So imagining the electron jelly bubble around each atom (and how thick/dense it is) is actually Useful !

Knew that would come in handy one day.

Pushing my ignorance a bit further ...

A Carbon atom with 4 bonds should be happy.

What we're discussing is that it's not always the case.

An adjacent bonded Oxygen is being greedy with the electrons, and the Carbon wants more, so is receptive to suitors bearing negatively charged Gifts ?

blogfast25 - 7-11-2015 at 07:12

Quote: Originally posted by aga  

A Carbon atom with 4 bonds should be happy.

What we're discussing is that it's not always the case.

An adjacent bonded Oxygen is being greedy with the electrons, and the Carbon wants more, so is receptive to suitors bearing negatively charged Gifts ?


Let's say that there are often energetically more favourable arrangements:

O
||
-C- (carbonyl) and -CH<sub>2</sub>- (methylene) both have 4 bonds on the C atom but the formation of a carbonyl group is more exoenergetic than the formation of a methylene group.

Now does this mean that compounds that contain methylene groups are all queuing up to be 'carbonylised'? No. As it happens that conversion has a high activation energy, so does not really happen spontaneously: thermodynamics does not equal kinetics!

[Edited on 8-11-2015 by blogfast25]

aga - 7-11-2015 at 12:13

Aha !

"Thermodynamics makes no pronouncement about Kinetics" leaps to mind.

AKA "Just cos it is possible does not make it happen"

blogfast25 - 7-11-2015 at 13:12

Quote: Originally posted by aga  
Aha !

"Thermodynamics makes no pronouncement about Kinetics" leaps to mind.

AKA "Just cos it is possible does not make it happen"


Exactly. Think butane lighter: no spark, no flame and yet Free Energy Change of butane combustion is plenty negative...

aga - 7-11-2015 at 14:50

OK. Exhausted my Butane lighter testing that out.

Can i bill the B&D Education Store for the replacement butane ?

blogfast25 - 7-11-2015 at 17:25

QQuiz Questions for next week: U2U me the answers!


Question 7: (for 10 points)

Estimate the Enthalpy of Formation (at 298 K) of fluoromethane (CH<sub>3</sub>F), from bond Enthalpies. Show your calculation.


Question 8: (for 20 points)

When 1-chloropropane is refluxed with a small stoichiometric excess of strong NaOH solution, 1-propanol and a solution of NaCl is obtained according to the following overall reaction equation:

propyl chloride hydrolysis.png - 3kB

Propose a simple reaction mechanism (use curly arrows to show the movements of electrons).


[Edited on 8-11-2015 by blogfast25]

Darkstar - 7-11-2015 at 23:52

Quote: Originally posted by aga  
Oh !

So the whole resonance thing is imagining (even calculating!) where the electron densities are on the molecule, especially the active bits, and the relative electrostatic charges that implies ?


There's a lot to resonance. It can help explain anything from why one molecule is more stable than another, to why certain reactions work with some molecules but fail with others. As I've said, having a good grasp on the idea behind resonance is extremely important in OC. There's a reason I keep stressing this concept and even dedicated an entire post to showing examples of resonance structures! So just to recap, in the previous quiz, we used resonance to explain:

a) why the proton on acetic acid's hydroxyl group is far easier to remove than the proton on ethanol's hydroxyl group; and

b) why the carbon of the acetate ion is no longer electrophilic enough to undergo further nucleophilic attack.

In the former (a), the negative charge on acetic acid's conjugate base (the acetate ion) ends up getting distributed over three different atoms due to electron delocalization, making it significantly weaker than the localized negative charge on ethanol's conjugate base (the ethoxide ion). Because of this, it is much easier to pull a hydrogen off of acetic acid than it is to pull one off of ethanol, as the attraction of the positively-charged hydrogen back to the negatively-charged oxygen is far weaker in acetic acid. Thus acetic acid is said to be a stronger acid than ethanol, while the ethoxide ion is said to be a stronger base than the acetate ion. In the latter (b), the partial-negative charge (-1/3) on the carboxyl-group carbon of the acetate ion protects it from any further nucleophilic attacks via electrostatic repulsion, making the reaction irreversible as long as the conditions remain basic enough to keep the acetate ion deprotonated.

To help further illustrate the impact that resonance can have on OC reactions, in my next post, I'm going to show you a rather interesting reaction that I have come to believe ONLY works due to resonance. I will contrast this reaction with another reaction that, due to a lack of resonance stabilization, ends up proceeding through a completely different mechanism, despite the fact that both reactions use similar substrates under the same conditions.

Darkstar - 8-11-2015 at 02:16

Quote: Originally posted by Darkstar  
To help further illustrate the impact that resonance can have on OC reactions, in my next post, I'm going to show you a rather interesting reaction that I have come to believe ONLY works due to resonance. I will contrast this reaction with another reaction that, due to a lack of resonance stabilization, ends up proceeding through a completely different mechanism, despite the fact that both reactions use similar substrates under the same conditions.


So a couple of months ago I made a post in an OC thread called HI+P reduction mechanism--a thread about the mechanism for the reduction of benzyl alcohols using hydroiodic acid (HI) and red phosphorous. If this sort of reaction sounds familiar to you, it's probably because it's commonly used by clandestine chemists to reduce l-ephedrine and d-pseudoephedrine (both benzyl alcohols) to d-methamphetamine (we'll talk about what the d- and l- prefixes mean when we get to chirality), and is the reason that iodine and red phosphorous are so closely watched in places like the US. This type of reduction is extremely well-known and was first discovered more than 100 years ago; however, the exact nature of its mechanism remains a mystery to this day, and there's still no unanimous consensus as to how it actually works. In the thread mentioned above, I proposed a mechanism for it (which you will see below) that is based on the various conclusions that I have found in the literature over the years.

But before I show you the reaction between HI and benzyl alcohols, I want to show you what a typical reaction between HI and an alcohol looks like. Since the benzyl alcohol in the example reaction is 1-phenylethanol, I will show you how HI normally reacts with secondary (and tertiary) alcohols. A secondary alcohol is one where the hydroxyl group is connected to a carbon that is further connected to two other carbons, by the way. Both of the reactions below involve heating a secondary alcohol with concentrated hydroiodic acid; however, despite being similar alcohols, they give completely different products. In one reaction, the hydroxyl group gets replaced by an iodine atom, which is what we would expect from treating an alcohol with a hydrohalic acid. In the other reaction, however, something unique and unexpected happens: the hydroxyl group gets replaced by a hydrogen atom instead! (what "reduction" means in organic chemistry)

So let's take a look at the usual reaction between secondary alcohols and HI. In this example, HI is reacted with propan-2-ol, also known as 2-propanol, isopropyl alcohol or simply "rubbing alcohol." Here, the hydroxyl group gets replaced with an iodo group via nucleophilic substitution, producing 2-iodopropane and a water molecule. This type of reaction is called an Sn1 reaction (unimolecular nucleophilic substitution) and is one of the most common types of reactions in OC. Here's the reaction and mechanism:

extra credit reaction1.bmp - 737kB

In the interest of keeping this post as short as possible, I will just briefly sum up what's happening:

Step 1 - The nucleophilic oxygen of the hydroxyl group donates a lone pair to the electrophilic hydrogen on hydronium. (remember that HI, when dissolved in water, is really in the form of a bunch of H3O<sup>+</sup> and I<sup>–</sup> ions) This breaks the old bond on hydronium and creates a new one on isopropyl alcohol, giving a charged oxonium ion.

Step 2 - The oxonium group on isopropyl alcohol takes the electrons with it and leaves as neutral water. This generates a positively-charged carbon (called a carbocation) at position-2, where the carbon has three bonds and is short an electron.

Step 3 to Step 4 - The nucleophilic iodide ion donates a lone pair to the electrophilic carbon, filling its vacant orbital and forming a new bond that neutralizes the charges.

This is the usual reaction between secondary alcohols and HI. Seems simple and straightforward enough, right? But when a benzyl alcohol is used in that same reaction, something strange happens. Like before, the hydroxyl group leaves and gets replaced by iodine (it's arguable whether or not it actually does, but for now we'll just assume it does); however, the molecule then reacts again and ends up getting reduced by iodide, giving the reduced alcohol, I2 and water. This reaction is a lot more complex than the ones we've looked at so far and involves radical intermediates, so don't worry if the mechanism makes absolutely zero sense after Step 2. A radical is when an atom has a single, unpaired valence electron, by the way. (which typically makes it extremely reactive) The arrows that look like they have a fishhook on the end mean that a single electron is being moved:

extra credit reaction2.bmp - 1.3MB

If you want more details regarding this mechanism, just let me know. Basically, the reaction proceeds as follows:

Step 1 - The hydroxyl group is protonated and leaves as water, giving a resonance-stabilized carbocation at the benzylic position.

Step 2 - Iodide attacks the carbocation and forms a new bond.

Step 3 - The weak carbon-iodine bond cleaves homolytically to give a resonance-stabilized benzylic radical and an iodine radical.

Step 4 - The benzylic radical is reduced via SET (single-electron transfer) from a nearby iodide ion to give a negatively-charged carbanion (essentially the opposite of a carbocation) and a second iodine radical.

Step 5 to Step 6 - The carbanion grabs a proton from hydronium to give the reduced alcohol and the two iodine radicals combine to give I2.

This is how I believe the reaction works, which is made possible through a combination of resonance-stabilization and iodine's unique properties due to its size. This mechanism is suggested in the literature and is further supported by the fact that most other alcohols that have a pi-system directly adjacent to the hydroxyl group get reduced as well, strongly suggesting some sort of resonance-stabilized radical intermediate. The benzylic radical and iodine radical seen in Step 4 are created via homolytic cleavage of the carbon-iodine bond, which is shown in Step 3. Homolytic cleavage is where the bond breaks and each atom takes one electron with it. Many times this requires a considerable amount of energy to do. But because of iodine's size, becoming a radical isn't too difficult for it; however, for a secondary carbon atom, normally it wouldn't want anything to do with that sort of thing. And as you can see in the reaction with isopropyl alcohol, there's no breakage of the C-I bond to generate radicals and thus no reduction.

But because of the unique stability of benzylic radicals due to the adjacent aromatic ring, the already-weak carbon-iodine bond becomes even weaker. (keep in mind that iodine-carbon bonds are already weak bonds to begin with) The benzylic radical is stabilized through resonance like this:

benzylic radical2.bmp - 390kB

And it is this resonance that makes the bond between the benzylic carbon and the iodine atom extremely easy to break homolytically and the unique reduction possible. The radical that would be generated in the first reaction with isopropyl alcohol if the C-I bond were cleaved homolytically would be a lot more unstable due to a lack of resonance, making the radical much more reactive and thus a lot more difficult to create. (still possible, though, just not as easy. in fact, at high enough temperatures, HI will even fully reduce carboxylic acids!)

blogfast25 - 8-11-2015 at 09:31

Revisiting the Wave Function

On stuff that doesn’t oscillate, determinism and the particle-wave duality

We’ve come this far in the QM/WM/QC webinar that the time has come to revisit (and hopefully further clarify) one the most fundamental concepts of this paradigm, the wave function ψ itself. Early on we started with the following definition:

Wave function definition.png - 30kB

We’ve since worked out the wave functions for a number of simple quantum systems and below we find the schematised wave functions of the ground state and first two excitations for a single particle in a 1 dimensional finite well (well potential is U<sub>0</sub>, instead of our usual notation V<sub>0</sub>;):

Finite potential well.png - 16kB

We’ve also seen that the wave functions ψ and associated energy values E are the eigensolutions of the time independent Schrödinger equation (here in 1D):

schrodinger equation.png - 6kB

At this point it’s necessary to dispel one of the most profoundly and frequently held misconceptions about Quantum wave functions that new students of the quantum world often entertain: the wave function as a ‘matter wave’. This pernicious, wrong and stubborn belief stems mainly from poorly digested interpretations of the de Broglie Hypothesis, skewed readings of the wave-particle duality and teaching analogies involving ‘standing matter waves’.

Looking at the wave functions of the finite well potential it is indeed very tempting to make the analogy with the standing waves that occur in a string of a musical instrument, with the ground state the lowest base frequency (that determines the 'pitch' of the note) and the excited states the so-called harmonics (that determine the 'timbre' of the note).

But nothing could be further removed from the truth: as one physicist famously remarked about quantum wave functions, “nothing oscillates there!” As we’ll shortly see, it would have been more complete to add, “apart from probabilities!

In the current and prevailing Born interpretation, the wave function is a probability function and the peaks and throughs probability amplitudes:

Born interpretation.png - 9kB

Here P(x) is the probability density distribution of the particle over its domain and the actual probability (of finding the particle) in an interval [x<sub>1</sub>, x<sub>2</sub>] to find is given by P( x<sub>2</sub>, x<sub>1</sub>;).

(Note: these relationships hold for Real and normalised wave functions)

The term ‘wave function’ is therefore somewhat misleading: unlike in water or sound waves, for ‘quantum waves’ there is no oscillating medium (water, air or any other elastic medium) and no propagation, so typical of water and sound waves.

Another misconception about QM that needs dispelling is: ‘QM is non-deterministic’, which largely stems from its seemingly probabilistic nature. But at least at the most basic level one can show this isn’t true.

We saw earlier on that the expectation value < x > can be calculated from a (Real and normalised) wave function as follows:

x expectation value.png - 5kB

By the expectation value < x > has to be understood the average position of the particle, taken over a very large number of measurements. For a single particle in an 1D infinite potential well of length L, we found that:

< x > = L/2

If using a detector we measure this value over a very large number of individual observations (or by observing a very large number of identical systems simultaneously) we always find < x > = L/2. Perfectly deterministic, in other words.

Finally, if quantum wave functions don't describe real waves, where does that leaves us with regards to the particle-wave duality we started this webinar from? Certainly moving subatomic particles and photons do sometimes behave like waves and sometimes like particles and we don't really have a good word for that. But to describe how moving particles show wave-like properties like diffraction, modern QM needs not to describe the particle as a ‘matter wave’. Instead the formalism of the particle as a probability wave is sufficient, so there is no contradiction there.

[Edited on 8-11-2015 by blogfast25]

Darkstar - 9-11-2015 at 01:48

I think we scared aga away.

Darkstar - 9-11-2015 at 04:27

Brief Introduction to Wedge-Dash Notation

When I have a little more time I'll do a proper lesson on chirality. For the time being, I've prepared a mini-lesson on wedge-dash notation. This will serve as a lead-in to enantiomers by familiarizing any young OC Padawans following this thread (*cough* aga *cough*) with the concept of representing three-dimensional molecules in only two dimensions.

While drawing molecules using the line-angle formula is undoubtedly the most practical way to draw molecules in OC, the problem is that molecules aren't actually two-dimensional in real life. For non-chiral molecules (we will cover chirality soon) this isn't really a problem; however, for molecules with one or more chiral centers, we need a way to represent which direction certain bonds are oriented in three-dimensional space because it makes a difference. With that said, I present to you the wedge-dash notation:

chirality intro1.bmp - 592kB

As you can see, the normal lines are oriented along the plane of the image; the wedged lines are pointing towards the viewer; the dashed lines are pointing away from the viewer. Compare the wedge-dash drawing of propane above to the ball-and-stick model of propane below:

Propane3D.bmp - 382kB

Notice that the bonds are oriented the exact same way as the bonds in the wedge-dash drawing suggest.

Extra Credit:

chirality intro2.bmp - 497kB

In the image above, I have drawn 2-chloropropane in the top box and (1-chloroethyl)benzene in the bottom box. In each case, I have drawn the molecule two different ways: one with the chloro group pointing towards the viewer, and one with the chloro group pointing away from the viewer.

For extra credit, explain why the two molecules in the top box are the exact same molecule, while the two molecules in the bottom box are not. What have I changed by adding a phenyl ring?

aga - 9-11-2015 at 05:36

Quote: Originally posted by Darkstar  
I think we scared aga away.

Not a chance !

There are a lot of pitchforks and red squiggly lines to absorb in your recent posts, which will take a little time.

blogfast25 - 9-11-2015 at 07:53

Answers to last week's QQuiz questions:


Quote: Originally posted by blogfast25  


Question 5: (for 10 points)

Predict the geometrical shape of the following molecules:

a) Xenon difluoride, XeF<sub>2</sub>.

b) The pentafluoride xenonate anion, XeF<sub>5</sub><sup>-</sup>.

Question 6: (for 10 points)

Dinitrogen difluoride (N<sub>2</sub>F<sub>2</sub>;) has two isomers, as shown below, left trans-N<sub>2</sub>F<sub>2</sub>, right cis-N<sub>2</sub>F<sub>2</sub>:


Briefly explain why two such isomers exist.




A. 5:

a. XeF<sub>2</sub>:

Both bonds are identical. Electrostatic repulsion between the bonding MOs is minimised if the molecule takes on a linear shape: F-Xe-F.

b. XeF<sub>5</sub><sup>-</sup>:

All bonds are identical and the single charge is delocalised. Electrostatic repulsion between the bonding MOs is minimised if the molecule takes on a trigonal bipyramidal shape as shown here.


A. 6:

Two of three valence electron of each N contribute to the double bond: two p<sub>x</sub> form a σ(pp) MO, two p<sub>y</sub> form a π(pp) MO, forming a torsionally rigid structure.

The F atoms can then bond to the remaining N valence electron on either side of the double bond. So two different structures, called 'cis' (same side) and 'trans' (one on each side) arise.

Additional note to A. 6:

Because of the abundance of double C bonds in organic chemistry, cis/trans isomerism is very common in OC, see below for example the structures for 2-butene:

trans and cis butene.png - 2kB

blogfast25 - 9-11-2015 at 08:14

Bond Enthalpies Erratum:

In the post on bond Enthalpies the following error was not previously spotted:


Quote: Originally posted by blogfast25  
Bond Enthalpies:


Estimating reaction Enthalpies from bond Enthalpies:

These values can be used to estimate the reaction enthalpies of simple reactions. Here, mainly for illustrative purposes, the estimated Enthalpy of formation of ethane:

2 C + 3 H<sub>2</sub> === > C<sub>2</sub>H<sub>6</sub>, ΔH<sub>f</sub><sup>298K</sup>

To find this value write this equation as a sum of ‘sub reactions’ using Hess’ Law:

1) 2 C === > C – C, yields - 368 kJ

2) 3 X [H<sub>2</sub> === > 6 H.], costs 3 X + 435 = + 1305 kJ

3) C – C + 6 H. === > H<sub>3</sub>C – CH<sub>3</sub>, yields 6 X – 414 = - 2484 kJ

Adding up: ΔH<sub>f</sub><sup>298K</sup> = - 368 + 1305 – 2484 = -1547 kJ per mol C<sub>2</sub>H<sub>6</sub>.


Although reaction Enthalpies can certainly be estimated by the method pointed out in the post, where that reaction is a formation reaction the Standard Enthalpy of Atomisation of Carbon, value +717 kJ/mol, needs also to be taken into account (this would also be true for the estimation of formation Enthalpies for non-carbon based covalent compounds, where the atomisation Enthalpy of the main element then needs to be accounted for).

In the case of ethane the calculated value of - 1547 kJ/mol needs to be adjusted by adding 2 X + 717 kJ/mol, which gives us a value of - 113 kJ/mol for the Standard Enthalpy of Formation of ethane. The generally listed value is - 84 kJ/mol.

Apologies.

[Edited on 9-11-2015 by blogfast25]

aga - 9-11-2015 at 14:31

Now i'm confused by the Questions.

Am i answering QM Q 7 with the revised or original methodology ?

Are there any OC Questions i've missed ?

Still a Whopping chunk of OC to re-read and digest some more (he said, hoping for mercy)

blogfast25 - 9-11-2015 at 15:02

Quote: Originally posted by aga  

Am i answering QM Q 7 with the revised or original methodology ?



As per our U2U use the 'revised'<sup>*</sup> method if you can. Scoring will reflect my error, so it should be an easy 10 smackaroons!



* Revised number = original + 717 kJ/mol. Easier than decimalisation!! :D

[Edited on 9-11-2015 by blogfast25]

Darkstar - 9-11-2015 at 15:14

Quote: Originally posted by aga  
Are there any OC Questions i've missed ?

Still a Whopping chunk of OC to re-read and digest some more (he said, hoping for mercy)


There's no rush; take your time. As far as OC questions, the only new one is at the bottom of this post. It's just an extra credit question to serve as a lead-in to our upcoming lesson on chirality and enantiomers, so don't worry about trying to figure it out unless you have time. It's just to get you thinking about visualizing molecules in three dimensions.

The other posts I made were just more on resonance. This post just more or less reiterates what I've already said before, and this post just talks about an interesting reaction between hydroiodic acid and benzyl alcohols (any alcohol with a pi-system directly adjacent to the hydroxyl group, really) that is likely a result of resonance stabilization. It's mostly just extracurricular reading since it's a bit more advanced than anything we've talked about up until now. I will say that the post does briefly introduce you to Sn1 reactions, however, which are extremely common in OC. It would probably be a good idea to at least read up on that part.

blogfast25 - 9-11-2015 at 18:00

Embroidering slightly more on bond Enthalpies:

In the post on bond Enthalpies I indicated that the listed values should be considered as averages, in the sense that the exact value of a bond Enthalpy depends on its molecular environment. The following link to a *.pdf illustrates that very well (besides also listing a far greater number of values):

https://labs.chem.ucsb.edu/zakarian/armen/11---bonddissociat...

(the values are bond dissociation enthalpies, so almost always positive)

Here's a sample of values for various C-C bonds in different molecules to illustrate the point of bond Enthalpy variability:

Bond Enthalpies 2.png - 22kB

aga - 10-11-2015 at 09:06

Quote:
Step 1 - The nucleophilic oxygen of the hydroxyl group donates a lone pair to the electrophilic hydrogen on hydronium.

Am i being Thick or does 'Lone Pair' mean TWO electrons ?

Is ONE electron being donated or TWO ?

blogfast25 - 10-11-2015 at 10:52

Quote: Originally posted by aga  
Quote:
Step 1 - The nucleophilic oxygen of the hydroxyl group donates a lone pair to the electrophilic hydrogen on hydronium.

Am i being Thick or does 'Lone Pair' mean TWO electrons ?

Is ONE electron being donated or TWO ?


I've drawn the two electron pairs (MOs) on the hydroxyl group:

Step 1.png - 9kB

It's the one marked green '1' that now attacks the proton on the hydronium (oxonium) and forms a bond with it. So it's not a question of 'donating ONE or TWO electrons': the lone, non-bonding MO 1 now becomes a bonding MO, bonding that second proton to the O atom.

Personally I would describe that step more easily as the protonation of that hydroxyl group by a H<sub>3</sub>O<sup>+</sup> cation (hydronium = oxonium): it's really a classic acid/base reaction:

R-OH(aq) + H<sub>3</sub>O<sup>+</sup>(aq) < === > R-OH<sub>2</sub><sup>+</sup>(aq) + H<sub>2</sub>O(l)

But in terms of movement of MOs, what I said above is what happens.

I prepared 2-chloropropane by that method but using ZnCl<sub>2</sub>(anh.) as a catalyst, not so long ago:

http://www.sciencemadness.org/talk/viewthread.php?tid=30180#...


[Edited on 10-11-2015 by blogfast25]

[Edited on 10-11-2015 by blogfast25]

aga - 10-11-2015 at 13:12

Sorry, still not Clear.

It's the term 'Lone Pair' that i'm struggling with.

Does that mean 'An orbital with between 0 and 2 electrons in it'
or does it mean 'An orbital with two electrons in it'
or does that mean 'An orbital with 0-2 electrons that is currently not involved in bonding' ?

The confusion arises from previous descriptions of electrons being swapped, and now we're into bonding orbitals where the electrons in it (or not) are somewhat abstracted.

blogfast25 - 10-11-2015 at 14:12

Quote: Originally posted by aga  
Sorry, still not Clear.

It's the term 'Lone Pair' that i'm struggling with.

Does that mean 'An orbital with between 0 and 2 electrons in it'
or does it mean 'An orbital with two electrons in it'
or does that mean 'An orbital with 0-2 electrons that is currently not involved in bonding' ?

The confusion arises from previous descriptions of electrons being swapped, and now we're into bonding orbitals where the electrons in it (or not) are somewhat abstracted.


In the context of Step 1:

An orbital with two electrons in it, that is currently not involved in bonding

But also in the context of Step 3:

Step 3.png - 5kB

The iodide ion has 4 lone electron pairs and uses one of them to bond with the carbonium ion with: one of the carbonium MOs is empty. The iodide acts as the Lewis base, the carbonium ion as the Lewis acid.

[Edited on 10-11-2015 by blogfast25]

aga - 10-11-2015 at 14:22

Thanks for that.

So 'Lone Pair' is a misnomer and should be interpreted as 'electron orbital' irrespective of it's apparent electron configuration, due to hybridisation ?

Sorry to labour this point, however i feel it is rather important to get the idea, seeing as it appears to be very important.

Edit:

If you would not mind, please explain it in terms of your recent illustrated example (the one with the OH, red arrow, H3O) etc)


[Edited on 10-11-2015 by aga]

blogfast25 - 10-11-2015 at 14:50

Quote: Originally posted by aga  
Thanks for that.

1) So 'Lone Pair' is a misnomer and should be interpreted as 'electron orbital' irrespective of it's apparent electron configuration, due to hybridisation ?



2) If you would not mind, please explain it in terms of your recent illustrated example (the one with the OH, red arrow, H3O) etc)




1) Yes! 'non-bonding molecular orbital' is probably the most accurate term, actually.

2) I prefer the other example:

Step 3.png - 5kB

The structure of the iodide anion is an sp<sup>3</sup> hybridised tetrahedron, all four MOs are full (2 electrons, Pauli conformant).

The electronic structure of the central, charge carrying C atom in the carbonium cation is:

* sp<sup>3</sup> hybridisation
* two of the sp<sup>3</sup> bond to two carbons
* one of the sp<sup>3</sup> bonds to one hydrogen
* one of the sp<sup>3</sup> is EMPTY!

Iodine then 'donates' one of its full sp<sup>3</sup> orbitals to the empty C sp<sup>3</sup> MO and a C-I covalent bond is formed. Et voila, everybody's happy.

aga too?


[Edited on 11-11-2015 by blogfast25]

blogfast25 - 11-11-2015 at 10:47

And here's a question for Darkstar and/or Annaandherdad.

I found it on a physics forum I'm a member of and got stuck trying to solve it.

Here's the question:

darkstar AAHD problem.png - 11kB

Another member part-answered the question as follows:

darkstar AAHD answer.png - 24kB

Because I'm not firm on my legs with Dirac notation, I wanted to calculate the energy expectation value using:

darkstar AAHD answer 2.png - 2kB

I first calculated the normalisation constant (the wave function is zero everywhere but in -a/2 < 0 < +a/2) and found it to be 10/√(61a<sup>5</sup>;).

But then it dawned on me we don't know V(x), so we don't know the Hamiltonian operator?

I'm stuck.

[Edited on 11-11-2015 by blogfast25]

aga - 11-11-2015 at 14:23

Yes ! Thanks for the detailed clarification.

Notes to self:-

1. Do not just count electrons.

2. Imagine the electron density 'clouds' to get a feel for where the charges are.

3. Lone Pair = unbonded orbital.

4. Use Electronegativity tables !
http://www.sciencemadness.org/talk/files.php?pid=421505&...

5. Electrophile wants more electrons
Nucleophile has too many.
N.B. RELATIVELY too many/few

6. Resonance is the dislocation of the bonding orbitals giving
more than 1 possible stable bonding configuration to a molecule

blogfast25 - 11-11-2015 at 14:38

In a nutshell, aga, in a nutshell...

annaandherdad - 11-11-2015 at 19:16

Blogfest, I got your message. The problem is not very well posed. The problem is probably trying to say that the potential is 0 for -a/2 < x < +a/2, and infinity outside that range. The energy eigenfunctions are

u_n(x) = sqrt(2/a) sin( 2n pi x/a)

and the energy eigenvalues are

E_n = (2/m) (n pi hbar /a)^2

However, the first question, what is the energy value you get on the first measurement, cannot be answered. The wave function is a linear combination of all of the energy eigenstates, so there are probabilities for measuring the different energy eigenvalues. You can't say which one will be measured, only the probabilities.

For the second question, the number of times you get the ground state is approximately the probability of being in the ground state times 1000.

The probability of being in the ground state is an integral that is a bit of a pain to do.

Darkstar - 12-11-2015 at 00:54

Quote: Originally posted by annaandherdad  
The problem is probably trying to say that the potential is 0 for -a/2 < x < +a/2, and infinity outside that range.


That's what I thought, too.

Quote: Originally posted by aga  
OK. Exhausted my Butane lighter testing that out.

Can i bill the B&D Education Store for the replacement butane ?


We give you a full scholarship to B&D University and you have the audacity to ask US to pay YOU for something that trivial? Well, I never!

blogfast25 - 12-11-2015 at 07:03

Quote: Originally posted by annaandherdad  
Blogfest, I got your message. The problem is not very well posed. The problem is probably trying to say that the potential is 0 for -a/2 < x < +a/2, and infinity outside that range. The energy eigenfunctions are

u_n(x) = sqrt(2/a) sin( 2n pi x/a)

and the energy eigenvalues are

E_n = (2/m) (n pi hbar /a)^2

However, the first question, what is the energy value you get on the first measurement, cannot be answered. The wave function is a linear combination of all of the energy eigenstates, so there are probabilities for measuring the different energy eigenvalues. You can't say which one will be measured, only the probabilities.

For the second question, the number of times you get the ground state is approximately the probability of being in the ground state times 1000.

The probability of being in the ground state is an integral that is a bit of a pain to do.


So there's no way to determine the amplitudes in the superposition? Fourier transform, anyone?

Can the superposition be be normalised? After all, the particle is bound...

Thanks!

Moderators: when will this site adopt MathJax (Latex rendering of math formulas) like below:

eigensolutions.png - 6kB

It's truly bizarre that on a forum about science in the 21<sup>st</sup> Century we still have to render the language of science with sub and sup tags...

[Edited on 12-11-2015 by blogfast25]

annaandherdad - 12-11-2015 at 07:56

Quote: Originally posted by blogfast25  

So there's no way to determine the amplitudes in the superposition? Fourier transform, anyone?

Can the superposition be be normalised? After all, the particle is bound...

Thanks!


Yes, the amplitudes of the superposition can be determiined, and their squares are the probabilities, but that doesn't tell you what energy value you will get on a measurement, only the probabilities of the different outcomes. You would only be able to predict exactly what energy would be measured if the system were in an energy eigenstate, which this one is not.

And yes, the superposition can be normalized.

I recall someone making the comment that the 1000 measurements had to be on independent systems, not repeated measurements on a single system, but I can't find the comment now. Anyway, that is correct, if you measure energy once, you leave the system in an energy eigenstate corresponding to the value of energy measured. Successive measurements of energy on the same system will yield the same value.

The probabilities given by quantum theory refer to measurements made on independent, identically prepared systems.

The wave function given (an inverted parabola) is qualitatively very similar to the ground state wave function (a sine function), so most of the probability will be in the ground state.

I think this problem requires too much algebra for the educational value, even it were properly stated.

blogfast25 - 12-11-2015 at 08:57

@AAHD:

Agreed with all your points. Thanks.

This is the thread:


http://physics.stackexchange.com/questions/217939/infinite-p...

aga - 12-11-2015 at 09:23

Quote: Originally posted by Darkstar  
For extra credit, explain why the two molecules in the top box are the exact same molecule, while the two molecules in the bottom box are not. What have I changed by adding a phenyl ring?


The molecules in the top box are the same thing rotated 180 degrees around the Y axis.

The ones in the bottom box are mirror images of each other.

I was about to triumphantly blurt 'steroisomer !' then had the sense to look that up.

The mirror-image thing makes them Enantiomers.

Edit:

If it's right, take the extra credits in liu of the butane ;)


[Edited on 12-11-2015 by aga]

Darkstar - 12-11-2015 at 10:11

Quote: Originally posted by aga  
The molecules in the top box are the same thing rotated 180 degrees around the Y axis.

The ones in the bottom box are mirror images of each other.

I was about to triumphantly blurt 'steroisomer !' then had the sense to look that up.

The mirror-image thing makes them Enantiomers.


Correct! They are mirror images of one another, called enantiomers. This is actually a very important concept, particularly in fields like medicine. Changing the spatial orientation of just a single atom or functional group in a biologically-active compound can sometimes drastically alter its effects. Welcome to the world of chirality.

Quote:
If it's right, take the extra credits in liu of the butane ;)


Well, I didn't want to bring this up, but I'm afraid you have forced my hand. There's a small matter of a room in one of the freshman dormitories that you completely trashed last weekend while drinking and doing unauthorized experiments with the undergrads. I've already received over two dozen complaints regarding broken glassware in the sink, chemical spills/splatters/spatters/puddles/sprays all over the walls, floors and counter tops, stains, holes in the carpet from what looks like a large acid spill, complaints of foul and obnoxious odors, hazardous waste that was improperly disposed of (halogenated compounds, toxic heavy metals etc), empty beer cans all over the place...

blogfast25 - 12-11-2015 at 10:22

Quote: Originally posted by Darkstar  
[...] while drinking and doing unauthorized experiments with the undergrads.


I heard it was experiments ON the undergrads. :D

aga - 12-11-2015 at 13:15

Quote: Originally posted by Darkstar  

Well, I didn't want to bring this up, but I'm afraid you have forced my hand. There's a small matter of a room in one of the freshman dormitories that you completely trashed last weekend while drinking and doing unauthorized experiments with the undergrads. I've already received over two dozen complaints regarding broken glassware in the sink, chemical spills/splatters/spatters/puddles/sprays all over the walls, floors and counter tops, stains, holes in the carpet from what looks like a large acid spill, complaints of foul and obnoxious odors, hazardous waste that was improperly disposed of (halogenated compounds, toxic heavy metals etc), empty beer cans all over the place...

As per the instructions of my Binary Logic Lecturer i will answer the allegations thusly :-
1110x1111111110x11 defintely 1

I take issue with the 'obnoxious odours' allegation.

They were merely natural odours that are much better Out than In.

The positioning of the Nose at such time as they become Out is a matter for the Smeller, not the Producer.

aga - 12-11-2015 at 15:22

Almost forgot the squigly diagram for QM question 8


q8.bmp - 100kB

Edit :

Seeing as we can't send images via U2U it might be easier to do the DarkStar thing and post the Answers in the thread, where images can be included.

OK with you bloggers ?

[Edited on 12-11-2015 by aga]

blogfast25 - 12-11-2015 at 15:30

Fur aga:

The normalisation procedure for the problem higher up:

Normalsation example.png - 25kB

Note that A has two roots, one positive, one negative, which represent the wave function's parity (remember those fellows gerate and ungerate?)

Now the wave function has been normalised we are certain that the probability density distribution P(x) is entirely correct. The wave function is now ready to extract information from.

Quote: Originally posted by aga  

Seeing as we can't send images via U2U it might be easier to do the DarkStar thing and post the Answers in the thread, where images can be included.

OK with you bloggers ?



As discussed, yes.


[Edited on 13-11-2015 by blogfast25]

aga - 12-11-2015 at 15:42

Ah feck.

i just saw that i got the H in the OH- attacking the 3-Carbon.

Should be the O ripping in there.

Oh the woes of MS Paint when applied to Organic Chemistry.

aga - 13-11-2015 at 13:49

Oh dear.

Looks like i scared the Professors away.

OK. I'll fix the dorm and sort out the polymerised undergrads.

It's just the chair legs and gravity holding them together.

Someone said 'stick and ball model' and i just got carried away.

They're only holding hands in mutual support - not to keep the structure intact or signify pi-bonds anymore.

Should all heal nicely for the Christmas break once the chairs are, er, extracted and returned to the canteen.

[Edited on 13-11-2015 by aga]

blogfast25 - 13-11-2015 at 15:28

Hydrolysis of 1-chloropropane by sodium hydroxide:

propyl chloride hydrolysis 2.png - 4kB

NaOH fully dissociates in water into sodium and hydroxide ions.

The nucleophilic (Lewis base) hydroxide anion attacks the polarised carbon atom. A new bond is formed, between the OH and the C atom.

The C-Cl MO folds back onto the Cl atom, forming a chloride ion as leaving group.


[Edited on 14-11-2015 by blogfast25]

Darkstar - 14-11-2015 at 03:47

Quote: Originally posted by aga  
i just saw that i got the H in the OH- attacking the 3-Carbon.


As far as nomenclature is concerned, the carbon that chlorine is bonded to is technically the 1-carbon. There are way too many nomenclature rules to attempt to explain them all in a brief post, but basically you number carbon chains based on what's attached to them, not left-to-right like you read. In 1-chloropropane (notice that it's not called 3-chloropropane!), you start on the carbon with the chloro group and count from there. On the other hand, if the chloro group were attached to the middle carbon, then you could start counting on either end since it wouldn't make a difference (either way, the carbon that the chloro group ends up on is the 2-carbon); however, you would NOT start counting on the second carbon because you always want the longest carbon chain as the parent chain. Carbon chains are named using a prefixparentsuffix structure (i.e. 1-chloropropane or 1-chloropropan-2-ol), where the parent chain is considered the "main" chain and everything else just "additions" to it.

In larger, more complex molecules with multiple functional groups, the order of precedence determines what gets prefixed and what gets suffixed. In other words, certain groups take priority over others. The most important group on a molecule gets suffixed at the end by itself, and all other groups get prefixed in alphabetical order in front of the parent chain. An exception to this are groups like chloro, bromo, iodo, fluoro and alkyl groups (methyl, ethyl, propyl, butyl etc), which never get suffixed regardless. In some cases, this priority will also determine which carbon you start counting on as well. Also, the prefix name for a functional group is usually different than the suffix name for it (i.e. when prefixed, hydroxyl groups are "hydroxy," but when suffixed, they are "ol").

But don't worry about trying to make sense of all of this right now. I will do a more in-depth lesson on IUPAC nomenclature in the future. Consider this extracurricular reading for the time being. Below are some examples I made for your viewing pleasure. Oh, and by the way, that last one is a sample of the kind of stuff I will be asking you to name on the homework questions!

nomenclature examples.bmp - 1MB

aga - 14-11-2015 at 09:12

Is that Fly Spray or Slug-knackerer ?

blogfast25 - 14-11-2015 at 09:27

Quote: Originally posted by aga  
Is that Fly Spray or Slug-knackerer ?


It lacks a methyl on C No 5 for that. :D

aga - 14-11-2015 at 12:08

Quote: Originally posted by Darkstar  
But don't worry about trying to make sense of all of this right now.

That's a Deal !

I suppose the vast array of OC entities makes even the Naming pretty complex. Sure looks that way.

aga - 15-11-2015 at 14:58

Please Sir(s), what does a Hartree-Fock equation do ?

Or a 'Virtual Orbital' mean ?

[Edited on 15-11-2015 by aga]

blogfast25 - 15-11-2015 at 17:43

Quote: Originally posted by aga  
Please Sir(s), what does a Hartree-Fock equation do ?

Or a 'Virtual Orbital' mean ?

[Edited on 15-11-2015 by aga]


Hartree-Fock lies at the heart of several approximation methods for multi-atom multi-electron systems (usually called molecules) ;)

For multi electronic systems the trusted Schrödinger equation:

schrodinger equation 2.png - 3kB

... has no analytical solutions because the inter-electron repulsive electrostatic forces cause Potential Energy (V) terms to occur in the Hamiltonian Operator, that aren't encountered in simpler one electron systems.

To find numerical solutions for these cases a number of approximative/iterative techniques have been developed. Hartree Fock is an important element in some of these methods.

Where did you find the term "Virtual Orbital"?


[Edited on 16-11-2015 by blogfast25]

blogfast25 - 15-11-2015 at 17:52

QQuiz Questions for the week will come tomorrow. Been a bit busy today.

[Edited on 16-11-2015 by blogfast25]

Metacelsus - 15-11-2015 at 19:52

Quote: Originally posted by aga  
Please Sir(s), what does a Hartree-Fock equation do ?


Basically, it ignores the instantaneous correlation of the wavefunctions, and instead treats wavefunctions like static charge distributions.

The Hamiltonian changes to an "effective Hamiltonian", which allows separation. The wavefunctions are solved for by an iterative process called the self-consistent field method.

By "virtual orbitals", do you mean Slater orbitals?

[Edited on 11-16-2015 by Cheddite Cheese]

aga - 16-11-2015 at 01:03

Quote:
Where did you find the term "Virtual Orbital"?


Same place i came across Hartree-Fock = York university Chem prospectus, QM module :-

http://www.york.ac.uk/chemistry/undergraduate/courses/option...

blogfast25 - 16-11-2015 at 07:21

Question 9: (for 10 points)

Calculate the Lattice Enthalpy of NaF(s) from the following data (all in STP conditions):

Enthalpy of Formation of NaF = - 575 kJ/mol
Enthalpy of vapourisation of Na = + 97.42 kJ/mol
Ionisation energy of Na (Na == > Na<sup>+</sup> + e) = + 495.8 kJ/mol
Enthalpy of bond dissociation of F<sub>2</sub> = + 159 kJ/mol
Electron affinity of F ( F + e == > F<sup>-</sup>;) = - 328 kJ/mol

Question 10: (for 20 points)

Many d-block elements form colourful aqueous complexes with ammonia. Explain why the following are colourless:

a) Cu(NH<sub>3</sub>;)<sub>4</sub><sup>+</sup>(aq)
b) Zn(NH<sub>3</sub>;)<sub>4</sub><sup>2+</sup>(aq)

Oh, and these are the last questions of the current QQuiz!


[Edited on 17-11-2015 by blogfast25]

blogfast25 - 16-11-2015 at 07:40

Answer to Question 7 (Question 8 has already been answered above):

The formation reaction equation of fluoromethane (methyl fluoride):

C(s) + 3/2 H<sub>2</sub>(g) + 1/2 F<sub>2</sub>(g) === > CH<sub>3</sub>F(g)

Enthalpies involved:

Enthalpy of atomisation of C = + 717 kJ

Breaking H2 bonds = 3/2 * 435 kJ

Breaking F2 bonds = 1/2 * 159 kJ

Three C-H bonds formed = 3 * (- 368) kJ

One C-F bond formed = - 536 kJ

Add it all up, gives - 191 kJ

No changes of state.

The estimated Enthalpy of fluoromethane in STP conditions is - 191 kJ/mol.

[Edited on 17-11-2015 by blogfast25]

aga - 16-11-2015 at 09:57

Woohoo ! I got that one ... hang on ... it says -191 ... DOH !

Q9 & 10 look hard. Real Hard.

Seeing as these are the last questions in this series, we need to organise the end-of-term party.

Who is bringing the booze & drugs this year ? Students or Masters ?

Darkstar - 16-11-2015 at 10:31

Quote: Originally posted by aga  
Who is bringing the booze & drugs this year ? Students or Masters ?


Why not just have students make it themselves for their final exam? It would test students on both theory as well as practice. Remember that reaction between hydroiodic acid and benzyl alcohols you learned about? Time to put that knowledge to good use.

aga - 16-11-2015 at 12:09

Erm, do you not remember the Wilkinson Affair when the Students did the drugs a couple of years back ?

Still thinks he's The Hyper Hampster as far as i know.

Not visited him for a long time.

Do you a deal : Students make 100% ethanol in 12 days, age it in 2 hours to a 40% brandy worthy of consumption.

Masters bring the drugs.

Smithers the Lab Tech will arrange the chicks as usual.

blogfast25 - 16-11-2015 at 19:51

Life on the edge of QM: Super-photons.

blogfast25 - 18-11-2015 at 18:20

Found this excellent QM resource below:

Not a QM course but an alphabetical index to Quantum Mechanical concepts and problems, with beautifully rendered mathematics:

http://www.physicspages.com/index-physics-quantum-mechanics/

From the same site:

Index to Calculus:

http://www.physicspages.com/index-mathematics-calculus/

Metacelsus - 20-11-2015 at 11:09

Question:

How exactly does Russel-Saunders coupling work? Also, how does it give rise to Hund's Rule?

aga - 20-11-2015 at 15:19

At the 11th hour, submitted my answers to the last QM questions of this series.

Not sure i did too well with Q 9 or Q 10, at all.

I can see the answers in the text of this thread, yet somehow cannot connect the dots.

Everyone else has probably sent in their answers already, so D-Day looms Large ...

Edit:

Whatever the Outcome, this text has been Amazingly educational, and will continue to be so for quite a time.

Tests are Good !

Looking back (to find answers) it becomes obvious which bits were not comprehended completely.

What do we want ?

More tests ! More QM ! More OC !

When do we want it ?

Now !

[Edited on 20-11-2015 by aga]

aga - 20-11-2015 at 15:38

Quote: Originally posted by Cheddite Cheese  
Question:

How exactly does Russel-Saunders coupling work? Also, how does it give rise to Hund's Rule?

Does this explain it ?

http://hyperphysics.phy-astr.gsu.edu/hbase/atomic/lcoup.html

Which one of Hund's 3 laws ?

The Multiplicity one ?

(yessir, i can google, i can google all night long ...)

annaandherdad - 20-11-2015 at 17:10

Quote: Originally posted by Cheddite Cheese  
Question:

How exactly does Russel-Saunders coupling work? Also, how does it give rise to Hund's Rule?


If you want to know exactly, it's a bit complicated. There is a very crude approximation to the structure of atoms, the central field approximation, which is based on a self-consistent picture in which every electron moves in a central force field (a rotationally symmetric potential, including both the "direct" and "exchange" potentials), which is determined by the states of all the electrons. The central field approximation is based on Hartree-Fock theory, and is described in my notes:

http://bohr.physics.berkeley.edu/classes/221/1112/notes/hart...

The central field approximation is too crude to give even a qualitative understanding of the low lying energy levels of multielectron atoms, although it does explain precisely the concept of the "electron configuration".

To get a qualitative and semi-quantitative understanding of the energy levels of atoms you need to do perturbation theory, based on the central field Hamiltonian as a zeroth order approximation. I started some notes on this,

http://bohr.physics.berkeley.edu/classes/221/1112/notes/atom...

but these are incomplete. The topic is continued in some hand-written notes I have,

http://bohr.physics.berkeley.edu/classes/221/0708/lectures/L...
http://bohr.physics.berkeley.edu/classes/221/0708/lectures/L...

The problem is to construct wave functions that have the right exact quantum numbers L and S of the electrostatic approximation to the exact Hamiltonian for the atom. This is carried out by some tabular methods that are explained in the notes I've linked. The main reference for my notes is Intermediate Quantum Mechanics by Bethe and Jackiw.

The tabular methods give you the various LS "multiplets" that can be created out of a given electron configuration. This is called Russel-Saunders coupling.

Hund's rules are semi-empirical rules that determine the order of the various LS multiplets within a given electron configuration.

In detail, it's a big topic, but if you want to know more detail about an aspect of it, let me know and I'll elaborate.

blogfast25 - 21-11-2015 at 09:46

Quote: Originally posted by aga  


What do we want ?

More tests ! More QM ! More OC !

When do we want it ?

Now !



Coming right up!

blogfast25 - 21-11-2015 at 10:11

Oh bugger it: it's the weekend, end of term and blahblah, so I'll divulge the answers to 9 and 10 a bit prematurely.

Quote: Originally posted by blogfast25  
Question 9: (for 10 points)

Calculate the Lattice Enthalpy of NaF(s) from the following data (all in STP conditions):

Enthalpy of Formation of NaF = - 575 kJ/mol
Enthalpy of vapourisation of Na = + 97.42 kJ/mol
Ionisation energy of Na (Na == > Na<sup>+</sup> + e) = + 495.8 kJ/mol
Enthalpy of bond dissociation of F<sub>2</sub> = + 159 kJ/mol
Electron affinity of F ( F + e == > F<sup>-</sup>;) = - 328 kJ/mol

Question 10: (for 20 points)

Many d-block elements form colourful aqueous complexes with ammonia. Explain why the following are colourless:

a) Cu(NH<sub>3</sub>;)<sub>4</sub><sup>+</sup>(aq)
b) Zn(NH<sub>3</sub>;)<sub>4</sub><sup>2+</sup>(aq)



A. 9:

The knowledge needed lies in this part of the course:

http://www.sciencemadness.org/talk/viewthread.php?tid=62973&...

The equation for the Enthalpy of Formation of NaF (SRD) thus becomes:

+ 97.42 + 495.8 + 1/2 159 - 328 + L = - 575

Thus the Lattice Enthalpy of NaF is:

L = - 919.2 kJ/mol, a suitably high (but negative) value for an ionic compound made of relatively small ions packed closely together. Remember that Lattice energy is mainly Coulombic potential energy.

A. 10:

The knowledge needed lies in this part of the course:

http://www.sciencemadness.org/talk/viewthread.php?tid=62973&...

For most colourful d-block element complexes the colour is explained by Ligand Field Theory. The ligands (ammonia molecules in this case) MO electron repulsion field causes a difference in energy level between the e and t group d orbitals. Electrons can now absorb VIS photons to transit between the two groups of d orbitals.

Now look at the specific cases of Cu(+1) and Zn(+2) complexes: from the electron configuration of both elements can be gleaned that the 3d orbitals of the ions are in fact full: 3d<sup>10</sup>. Thus these electrons cannot transit because all 3d orbitals are full.

By contrast, Cu(+2) shows a 3d<sup>9</sup> configuration, so one 3d orbital is now half-full. Another 3d electron can now fill that gap. Cu(+2) ammonia and aqua complexes are strongly blue.


[Edited on 21-11-2015 by blogfast25]

Darkstar - 21-11-2015 at 10:58

I'll post some more OC lessons soon. It's been a busy week for me!

aga - 21-11-2015 at 11:09

DOH! DOH! and DOH!

So the full d orbitals are the reason for the lack of colour.

DOH!

aga - 21-11-2015 at 12:09

Quote: Originally posted by Darkstar  
I'll post some more OC lessons soon.

You'd better, or we'll all dehydroxyfilicate, spontaneously.

blogfast25 - 21-11-2015 at 13:54

Here's a quick one, off the cuff:

Explain why Yttrium (Y, Z = 39) compounds are colourless.

blogfast25 - 21-11-2015 at 14:15

And one for the road.

Most alcohols react neutral with pure water. Solutions of phenol in water are slightly acidic.

Phenol.png - 1kB

Explain this difference from a structural point of view.

[Edited on 21-11-2015 by blogfast25]

aga - 21-11-2015 at 14:26

It's cos they do not absorb/emit any photons in the 390nm to 700nm range, having carelessly donated their sole 4d electron so some charity case or other.

Some elements are dangers to themselves.

blogfast25 - 21-11-2015 at 14:53

Draw all isomers of this alcohol:

pentanol.png - 2kB

aga - 21-11-2015 at 15:00

In MS Paint ?

You're kidding.

Numbering the carbons left to right, the OH can go on the 2, 3 or 4 postions.

Not entirely sure why it can't go on 1 or 5, but it just feels like it can't.

blogfast25 - 21-11-2015 at 15:12

Quote: Originally posted by aga  
It's cos they do not absorb/emit any photons in the 390nm to 700nm range, having carelessly donated their sole 4d electron so some charity case or other.

Some elements are dangers to themselves.


Correct. No d-electrons, no transitions. 10 (full set!) d electrons, no transmissions either... simples really! So 0 and 10: no colours.

Important corollary:

f block elements are also often colourful for similar reasons.

The 4 d orbitals also split into two groups according to orientation and thus slightly different energy levels, in a ligand field. An incompletely filled 4d orbital thus also gives rise to transitions with emissions in the VIS part of the electromagnetic spectrum. Most (+3) lanthanides show colour, as do many actinides due to part-filled 5d orbitals.


[Edited on 22-11-2015 by blogfast25]

blogfast25 - 21-11-2015 at 15:17

Quote: Originally posted by aga  
In MS Paint ?

You're kidding.

Numbering the carbons left to right, the OH can go on the 2, 3 or 4 postions.

Not entirely sure why it can't go on 1 or 5, but it just feels like it can't.


It can be done in MS Paint-in-the arse because the shapes are simple.

It can go from 1 to 5: it's a pentanol. No naming was requested (Darkstar will do the IUPAC incantations later on, I think).

[Edited on 21-11-2015 by blogfast25]

aga - 22-11-2015 at 09:00

Quote: Originally posted by blogfast25  
Draw all isomers of this alcohol


isomers.png - 36kB

'owzat ?

Metacelsus - 22-11-2015 at 09:14

There are other isomers. Keep on trying.

(And some of the compounds have two enantiomers, but I'm not sure we're counting those.)

[Edited on 11-22-2015 by Cheddite Cheese]

blogfast25 - 22-11-2015 at 09:18

Not bad but there's two more.

Think about methylbutane...

Try also the phenol question. Then I'll QC explain strong and weak acids, very briefly...

blogfast25 - 22-11-2015 at 09:21

Quote: Originally posted by Cheddite Cheese  
There are other isomers. Keep on trying.

(And some of the compounds have two enantiomers, but I'm not sure we're counting those.)

[Edited on 11-22-2015 by Cheddite Cheese]


Sorry, CC, the question specified alcohols. Sorry if that wasn't clear.

I wasn't counting enantiomers but point them out, by all means.


[Edited on 22-11-2015 by blogfast25]

Metacelsus - 22-11-2015 at 09:26

Aren't there three (not two) more alcohols?

[Edited on 11-22-2015 by Cheddite Cheese]

blogfast25 - 22-11-2015 at 09:55

Quote: Originally posted by Cheddite Cheese  
Aren't there three (not two) more alcohols?

[Edited on 11-22-2015 by Cheddite Cheese]


You're correct: there are three more.

And two out of the six are enantiomers, making a total of 8.

[Edited on 22-11-2015 by blogfast25]

aga - 22-11-2015 at 11:22

More quiggles



isomers.png - 57kB

blogfast25 - 22-11-2015 at 11:52

aga:

4, 5, 6 etc are conformations, not really isomers in the intended sense of the word.

Hint: try starting from 2-methyl butane, add the OH and see what you get. A C5 alcohol does not require all Cs to be 'inline', branching can still make a 'C5 ol'.

[Edited on 22-11-2015 by blogfast25]

aga - 22-11-2015 at 11:54

More squiggles !!! OK.

isomers.png - 52kB


Hmm

Looks like Foothark to me.

[Edited on 22-11-2015 by aga]


[Edited on 22-11-2015 by aga]

blogfast25 - 22-11-2015 at 13:36

The first three you started off with were fine and the very first one (pentan-2-ol) is an enantiomer. The others again are conformations.

Here are the other three, with the second one being an enantiomer:

pentanols.png - 4kB

The way to find enantiomers is to look for carbon atoms that are bonded to four different groups.

Take the middle one in my drawing. The C-atom bonded to the OH group is also bonded to a methyl group, an isopropyl group and finally a hydrogen atom. Four different species.

Such a molecule, when depicted in 3D always has a mirror image molecule. Together the pair are a pair of enantiomers.

[Edited on 23-11-2015 by blogfast25]

aga - 22-11-2015 at 13:54

The Volume of information in this thread is a little overwhelming for a part-time amateur.

It is amazingly condensed and informative, and well explained, therefore accessible.

In order to understand it more thoroughly (as it deserves) i've decided to cut down on the beer intake and devote the extra brain-function to learning this material better.

This a dangerous undertaking : the last time i stopped drinking altogether is when i took up Chemistry, so there's no knowing What may happen.

[Edited on 22-11-2015 by aga]

Metacelsus - 22-11-2015 at 14:57

Neopentanol (aka 2,2-dimethylpropanol) hasn't been mentioned yet. That was the third one I was thinking of.

Screen Shot 2015-11-22 at 4.56.22 PM.png - 9kB

[Edited on 11-22-2015 by Cheddite Cheese]

blogfast25 - 22-11-2015 at 15:10

More information overload fer aga (all resistance is futile).

Enantiomers are also referred to as optical isomers because the two isomers (sometimes called S and R) change the polarisation orientation of plane-polarised light in opposite ways. Guess what? That's a QM effect too!

If you can stomach it, here's a decent explanation on how that works.

[Edited on 23-11-2015 by blogfast25]

blogfast25 - 22-11-2015 at 15:15

Quote: Originally posted by Cheddite Cheese  
Neopentanol (aka 2,2-dimethylpropanol) hasn't been mentioned yet. That was the third one I was thinking of.



Yup. Good one. Should have looked at propane-based pentanols too.

blogfast25 - 22-11-2015 at 18:19

And while we're still talking about alcohols, don't forget this question about phenol... It's very informative, I promise!

[Edited on 23-11-2015 by blogfast25]

blogfast25 - 23-11-2015 at 08:24

A little Ode to Aga...

We've decided not to publish the scores of the QQuiz' participants but it needs to be said that aga's score constituted a Big Fat Pass (albeit perhaps not 'cum laude') Considering that this imperfect and incomplete course crams in material in about 4 months that at GCSE or A-level would take a full school year, that score is for a near-complete novice in fact an achievement. His 'dogged pursuit of knowledge' is equally commendable.

So three cheers for aga! The prize (a dented pingpong ball and a 'I heart Suzie Q' sticker manually modified to read 'I heart QM') is in the post!

Light relief:

I found this nice little page on alkane isomers that lists these isomers up to tetradecane (1858 isomers for that one!) The short introductory video is also worth watching.

[Edited on 23-11-2015 by blogfast25]

aga - 23-11-2015 at 08:39

A Pass ?! Erm, are you absolutely sure ? Astonishing.

The glory is all yours blogfast, not forgetting Darkstar's superb OC summary.


blogfast25 - 23-11-2015 at 11:20


Aga: Ok then. Now let's look at the following (fairly big) chunk.

Express recap of Brønsted–Lowry (BL) acid base theory:

Bronsted acid base.png - 11kB

The indexed Cs in the K equations represent actual concentrations (assuming a dilute solution), not what it says on the bottle (for “0.1 M acetic acid” e.g. the C<sub>HA</sub> is always (slightly) lower than the nominal concentration of 0.1 M).

It has to be noted also that A<sup>-</sup>, the so-called conjugated base of HA, is in fact also a weak base. See the alkaline solutions of acetates, carbonates, bicarbonates etc.

If B is the conjugated base of HA then we can even derive (derivation not given here) that:

pK<sub>a</sub> + pK<sub>b</sub> = 14

K and pK values thus offer an objective criterion by which we can rank acids and bases from strong to weak.

Stabilisation of Phenoxide anion:

The deprotonation reaction of phenol in water is of course:

phenol as acid.png - 13kB

We know that non-aromatic alcohols react neither acidly nor basic, so what prompts phenol to be such a (very!) weak acid? The answer lies in the stabilisation of the phenoxide anion due to de-localisation of the negative charge. The full explanation can be found here. A de-localised negative charge is less attractive to protons.

Why strong acids are strong acids:

Let’s recall the case of HBr:

HBr dipole.png - 1kB

The H-Br bond is highly polarised due to the large difference in electronegativity between Br and H. In the presence of a willing proton receptor like water the deprotonation reaction of HBr runs to completion (equilibrium fully to the right). HBr solutions contain almost no free, undissociated HBr, containing only bromide and oxonium ions (and solvent water).

Cross-paradigm agreement:

In a decent country with reasonable laws and rules these regulations apply from north to south and east to west.

The same is true of a good scientific paradigm (set of consistent theories). And here we have quite an interesting case of correspondence between Brønsted–Lowry acid base theory and its complementary Lewis acid base theory.

Consider the case of an interesting chloro-complex, called hexachlorostannate:

Hexachloro stannate.png - 4kB

Without going into MO structure (which is simple) the black orbitals lie in the xy-plane, the red ones align with the z-axis. The structure is in accordance with a typical EX<sub>6</sub> molecule, an octahedron.

Cl<sup>-</sup> can be seen as the conjugated base of HCl, an acid slightly stronger even than HBr. By definition higher up Cl<sup>-</sup> must be an extremely weak base, as of course evidenced by the neutrality of NaCl solutions.

So if Cl<sup>-</sup> is a weak BL base, is it also a weak Lewis base? Actually, yes.

Firstly, is it a Lewis base at all? Yes. In hexachlorostannate, the bonds are formed by six Cl<sup>-</sup> donating one of its four lone pair electron to empty hybrids (sp<sup>3</sup>d<sup>2</sup>, to be precise). That’s the essence of a Lewis base: an electron pair donor.

Hexachlorostannate salts (ammonium and alkali metals) can be crystallised easily from water. But they’re not hugely stable:

1. Adding a little ammonia to their solutions immediately precipitates Sn(OH)4. Adding stronger alkali leads to formation of stannate: Sn(OH)<sub>6</sub><sup>2-</sup>. This due to ligand exchange: hydroxide ions are a far stronger Lewis base.

2. Heating K2SnCl6 to above the BP of SnCl4 causes the latter to distil off.

Many other chloro-complexes are far less stable: tetrachlorocuprates and tetrachloroferrates exist in solution but cannot be isolated from them (at least not by simple evaporation of solvent), KAlCl4 does exist but can’t be recrystallised from water. All three are subject to ligand exchange with hydroxide ions (ammonia solution, for instance).

So Cl<sup>-</sup> is a very weak BL base and isn’t all too great as a Lewis base either.


[Edited on 24-11-2015 by blogfast25]

aga - 23-11-2015 at 12:20

Quote: Originally posted by blogfast25  

...

Without going into MO structure (which is simple) the black orbitals lie in the xy-plane, the red ones align with the z-axis. The structure is in accordance with a typical EX<sub>6</sub> molecule.

...

In hexachlorostannate, the bonds are formed by six Cl<sup>-</sup> donating one of its four lone pair electron to empty hybrids (sp<sup>3</sup>d<sup>2</sup>, to be precise).

Being a mere 'Pass' retard, the first thing i did was go look at the electron configuration of Sn. It didn't help.

Not having re-read and fully comprehended the whole thing yet, it appears that i missed something fundamental in how the hybridisation happens (maybe other key bits too).

A step-by-step explanation of what the electrons are doing in the formation of this ion would be very very helpful.

Pretty pretty please.

blogfast25 - 23-11-2015 at 13:21

Quote: Originally posted by aga  


A step-by-step explanation of what the electrons are doing in the formation of this ion would be very very helpful.

Pretty pretty please.


Alright, but only 'cos you ask so nice...

Sn ground state: [Kr] 4d<sup>10</sup>5s<sup>2</sup>5p<sup>2</sup>.

The central Sn ion in the hexachlorostannate ion is in oxidation state +4, so it has lost these four 5s<sup>2</sup>5p<sup>2</sup> valence electrons, these orbitals are now empty.

The empty 5s, three 5p (one p<sub>x</sub>, one p<sub>y</sub> and one p<sub>z</sub>;) and two 5d orbitals (also empty) now hybridise into 6 sp<sup>3</sup>d<sup>2</sup> orbitals, all empty.

The chloride ions are soft Lewis bases because they each have 4 lone electron pairs: electron configuration of Cl<sup>-</sup> is [Ne] 3s<sup>2</sup>3p<sub>x</sub><sup>2</sup>3p<sub>y</sub><sup>2</sup>3p<sub>z</sub&g t;<sup>2</sup>.

Each chloride ion now donates one of its lone electron pairs into one of the 6 empty sp<sup>3</sup>d<sup>2</sup>. Each molecular orbital resulting from this is full. These MOs are essentially σ MOs.

Clearish, squire?

[Edited on 23-11-2015 by blogfast25]

 Pages:  1  ..  3    5    7  ..  9