1. Po raz pierwszy odwiedzasz EDU. LEARN

    Odwiedzasz EDU.LEARN

    Najlepszym sposobem na naukę języka jest jego używanie. W EDU.LEARN znajdziesz interesujące teksty i videa, które dadzą Ci taką właśnie możliwość. Nie przejmuj się - nasze filmiki mają napisy, dzięki którym lepiej je zrozumiesz. Dodatkowo, po kliknięciu na każde słówko, otrzymasz jego tłumaczenie oraz prawidłową wymowę.

    Nie, dziękuję
  2. Mini lekcje

    Podczas nauki języka bardzo ważny jest kontekst. Zdjęcia, przykłady użycia, dialogi, nagrania dźwiękowe - wszystko to pomaga Ci zrozumieć i zapamiętać nowe słowa i wyrażenia. Dlatego stworzyliśmy Mini lekcje. Są to krótkie lekcje, zawierające kontekstowe slajdy, które zwiększą efektywność Twojej nauki. Są cztery typy Mini lekcji - Gramatyka, Dialogi, Słówka i Obrazki.

    Dalej
  3. Wideo

    Ćwicz język obcy oglądając ciekawe filmiki. Wybierz temat, który Cię interesuje oraz poziom trudności, a następnie kliknij na filmik. Nie martw się, obok każdego z nich są napisy. A może wcale nie będą Ci one potrzebne? Spróbuj!

    Dalej
  4. Teksty

    Czytaj ciekawe artykuły, z których nauczysz się nowych słówek i dowiesz więcej o rzeczach, które Cię interesują. Podobnie jak z filmikami, możesz wybrać temat oraz poziom trudności, a następnie kliknąć na wybrany artykuł. Nasz interaktywny słownik pomoże Ci zrozumieć nawet trudne teksty, a kontekst ułatwi zapamiętanie słówek. Dodatkowo, każdy artykuł może być przeczytany na głos przez wirtualnego lektora, dzięki czemu ćwiczysz słuchanie i wymowę!

    Dalej
  5. Słowa

    Tutaj możesz znaleźć swoją listę "Moje słówka", czyli funkcję wyszukiwania słówek - a wkrótce także słownik tematyczny. Do listy "Moje słówka" możesz dodawać słowa z sekcji Videa i Teksty. Każde z słówek dodanych do listy możesz powtórzyć później w jednym z naszych ćwiczeń. Dodatkowo, zawsze możesz iść do swojej listy i sprawdzić znaczenie, wymowę oraz użycie słówka w zdaniu. Użyj naszej wyszukiwarki słówek w części "Słownictwo", aby znaleźć słowa w naszej bazie.

    Dalej
  6. Lista tekstów

    Ta lista tekstów pojawia się po kliknięciu na "Teksty". Wybierz poziom trudności oraz temat, a następnie artykuł, który Cię interesuje. Kiedy już zostaniesz do niego przekierowany, kliknij na "Play", jeśli chcesz, aby został on odczytany przez wirtualnego lektora. W ten sposób ćwiczysz umiejętność słuchania. Niektóre z tekstów są szczególnie interesujące - mają one odznakę w prawym górnym rogu. Koniecznie je przeczytaj!

    Dalej
  7. Lista Video

    Ta lista filmików pojawia się po kliknięciu na "Video". Podobnie jak w przypadku Tekstów, najpierw wybierz temat, który Cię interesuje oraz poziom trudności, a następnie kliknij na wybrane video. Te z odznaką w prawym górnym rogu są szczególnie interesujące - nie przegap ich!

    Dalej
  8. Dziękujemy za skorzystanie z przewodnika!

    Teraz już znasz wszystkie funkcje EDU.LEARN! Przygotowaliśmy do Ciebie wiele artykułów, filmików oraz mini lekcji - na pewno znajdziesz coś, co Cię zainteresuje!

    Teraz zapraszamy Cię do zarejestrowania się i odkrycia wszystkich możliwości portalu.

    Dziękuję, wrócę później
  9. Lista Pomocy

    Potrzebujesz z czymś pomocy? Sprawdź naszą listę poniżej:
    Nie, dziękuję

Już 62 479 użytkowników uczy się języków obcych z Edustation.

Możesz zarejestrować się już dziś i odebrać bonus w postaci 10 monet.

Jeżeli chcesz się dowiedzieć więcej o naszym portalu - kliknij tutaj

Jeszcze nie teraz

lub

Poziom:

Wszystkie

Nie masz konta?

4. Freshman Organic Chemistry: Coping with Smallness and Scanning Probe Microscopy


Poziom:

Temat: Edukacja

Prof: Okay, so we saw last time that Lewis
wasn't so dumb.
He knew about Earnshaw's theorem, but he still thought
there was an octet of electrons around the nucleus.
How could he think such a thing?
Because if you have inverse-square force laws you
can't have static positions of charged things,
unless they're right on top of one another or blown apart.
So what did he think?
Student: It's not an inverse-square.
Prof: That it's not an inverse-square force law,
that Coulomb's law breaks down when you get to a very small
scale.
And J.J. Thomson thought the same thing, and wrote here these
two terms here.
The first one is Coulomb's law indeed, but there's this other
thing, c/r.
So as long as r is very big you don't compare -- you
don't care about that term, as long as it's very big
compared to c.
And c is a distance that's something like the size
of atoms.
So once r -- once the distance of the things that are
interacting with one another, the electron and the nucleus
say, once they get near the distance c,
then this c/r thing becomes significant,
if r becomes very small, and it even changes the sign.
So what was attractive becomes repulsive.
So that would do okay then for a structure.
That was in 1923.
In 1926, just three years later, quantum mechanics came
along and showed that Coulomb's law was just fine,
nothing wrong with Coulomb's law.
It goes to much, much, much, 10^20th,
smaller distances than the size of atoms.
Coulomb's law still works.
What was wrong was the way they treated kinetic energy,
because kinetic energy quantum mechanics reformulates.
So it gives you electron clouds.
And so what's a cloud is not the positive charge -- remember,
J.J. Thomson had a positive sphere of electricity in which
he embedded the electrons -- what's a cloud is the
electrons, and nuclei are put in it,
to make molecules.
So it really is -- he was right about plum-puddings,
he just had the charges backwards.
Not many people give him credit for that.
So cubic octets of Lewis and ad hoc electrostatic
force laws, like this c/r term in
there, soon disappeared from
conventional chemistry and physics;
immediately in fact, I mean within a month say.
But the idea of shared-pairs, and lone-pairs,
that Lewis came up with, remained useful tools for
discussing structure and bonding.
So we still do it, and that's why you did problems
today that had to do with that.
Yes Niko?
Student: Yes.
Can you explain the kinetic energy theorem again?
Prof: Yes, but I won't do it for a week
still, because it takes a little while.
Okay, so Earnshaw said there's no structure of minimum energy
for point charges -- that is, no classical one --
but if you come to quantum mechanics and fiddle around with
kinetic energy, then you can get it.
But despite Earnshaw, might Lewis have been right?
Might there still be shared-pair bonds and
lone-pairs?
What are bonds?
And that's what we're going to try to find out.
And how do you know, or how do you know?
How are we going to know whether there are electron pairs
that are shared between atoms and lone-pairs on atoms?
How can we find out?
Student: Experiment.
Prof: Right.
You've got to do an experiment.
So what kind of experiment?
Now these experiments are tough.
Why are the experiments tough?
Student: It's so small.
Student: It's really small.
Prof: Because things are -- well we could see,
or we could feel whether these electron pairs are there.
But there's a problem, as you say, that it's
inconceivably small, the thing we're looking for,
to see or to feel.
Okay?
But is it really inconceivable how small they are?
When I was in your place everybody thought yes,
they're inconceivably small, we'll forget it.
Right?
But there are more recent techniques that suggest that you
can really see things that small;
at least you can think about them clearly.
Now, the first idea of seeing them came from this book here.
Actually this book was published -- this is a Third
Edition of the book.
The First Edition was 1909 and the Second was 1919,
and this is the Third Edition, which is still in print.
You can go on the web and buy this baby,
Occult Chemistry, A Series of Clairvoyant
Observations on the Chemical Elements,
by Annie Besant and Charles Leadbeater.
So these are the occult chemists who started practicing
their trade in 1895.
"Bishop" Charles Webster Leadbeater;
he was a bishop of a church he founded that had -- he was the
head bishop; there were like 100 bishops and
500 members of the religion.
This is Annie Besant, who is one of the most
interesting people in all of history.
They closed the stock exchange in India the day she died,
for example.
This is in her -- the outfit she designed to lead the Boy
Scouts in India.
And she also proposed marriage to George Bernard Shaw.
There are all sorts of interesting things about Annie
Besant.
But one of the things we're interested in is that she could
see atoms.
And then their helper, who became leader of the
operation later on, was Jinarajadasa.
Okay, so this is a page from this book,
from the paper that became the book in 1895,
which shows what they could see for hydrogen,
oxygen and nitrogen.
Now there are a number of levels here.
The very lowest phase is solid, H O N, then liquid,
then gas.
So this is what, when they went into sort of a
trance, they could see.
This is the gas.
But then there's Ethereal 4, Ethereal 3, Ethereal 2 and
Ethereal 1 phases as well, with higher and higher
resolution.
You can see these.
So these, for example, these little dots that you see
here are what get blown up here, and the little things inside
this are what get blown up here, and the things inside this get
blown up here, and the things inside this are
this; and that's the same for all
atoms, that's the fundamental unit.
Okay, so here's hydrogen at the gas resolution level.
So there are wheels in wheels in wheels here.
And here's what oxygen looks like;
it's these little helices.
And now that's the thing you finally get to,
the thing called anu.
And that they actually ripped off from this book,
The Principles of Color, which was published in America
in 1878 by Edwin D. Babbitt.
And here's his picture of this thing.
And lo and behold, that's what they saw as the
fundamental constituent of matter.
It's fun to look at in detail.
You can look at it at the Web, if you want to.
It also has pictures of the Angel of Innocence here,
and the psycho-magnetic curves surrounding your head.
Okay, now lo and behold the real confirmation of this thing
was that if you counted up the anu in the different
elements, there were eighteen anu
in hydrogen, 290 in oxygen and 261 in
nitrogen, and if you divide that by eighteen,
you get the atomic weights, Q.E.D. Okay,
so they have pictures in this book of a bunch of the atoms.
This is helium, which has seventy-two
anu.
This is lithium with 127 anu.
Not the bottom part; this is a model and this is the
wooden base it sits on.
Here's iron, 1008 anu.
Here's neon, 360 anu.
Now look at neon, and that is a 4f orbital
electron density, calculated by quantum
mechanics.
Now the question is, why would you believe me,
or believe a textbook, or believe a quantum
mechanician, and not believe them?
Because they said they saw it too.
Okay?
Well it takes time.
We've talked about this before, the basis for scientific
belief: evidence, you always cross-examine an
assertion; logic; and taste, but taste matures
with experience and at the beginning you don't know which
people to believe and which not to believe.
It's a problem in politics too.
Okay, there's sodium, 418.
They not only said they saw atoms, they also saw molecules.
For example, here is sodium carbonate.
What do you see in there?
Two sodiums, right?
Okay?
In the back are pieces of a carbon atom.
What else do you see?
Students: Oxygen.
Prof: Ah, there are three oxygen helices
wrapped around the sodium and the carbon.
And they wrote: "Note that this triangular
arrangement of O_3 has just been deduced by Bragg from his X-ray
analysis of Calcite."
We'll talk about Bragg and what he was doing next time.
But they have experimental evidence that supports what
they're reporting, independent experimental
evidence.
Here's their model of benzene, and they say about that,
"Each of the three valencies of each Carbon are
satisfied by Hydrogen, and the fourth valency,
which some have postulated as going to the interior of the
molecule, does actually do so."
In fact, Annie Besant was the first woman to get a degree in
Chemistry at the University of London;
but that's another story.
Her teacher was the guy who translated Das Kapital
into English, and had a lot of other
nefarious things that he did.
Question six will have to do with that.
Okay, so here's benzene and its resonance.
These were structures that were first early proposed;
we'll talk about that later on.
But some people didn't like this idea of going back and
forth between two things, and they thought that you just
have the fourth valence of every carbon going into the middle,
all satisfying one another.
So this was what they said that their model confirmed,
that the fourth valence of every carbon goes into the
middle.
Okay, needless to say most people nowadays don't think too
much of occult chemistry, even if you can still buy the
book online.
This portal you came through this morning.
It says "Erected 1921".
But here's the same place, in 1923, when the building was
dedicated, and you notice those plaques weren't carved yet.
It didn't say erected 1921.
They were still finishing the building at the time the
American Chemical Society had their national meeting here to
celebrate the biggest academic building for chemistry in the
world.
Okay?
So the slates were clean.
Now if you go down into the crowd here, you see some
interesting people.
For example, there's G.N. Lewis with his
Philippines' cigar in his hand.
He always had a cigar; he smoked seventeen cigars a
day.
>
Groan.
Okay, but they were trying to decide how to decorate all these
plaques around the building, and they wanted to put the
names of distinguished American chemists.
Had they done that, it would be pretty much a
disaster because the people who were thought to be distinguished
American chemists in those days were not so hot actually in
retrospect.
But they chose a wonderful group, and the idea came from
Giuseppe Bruni from Bologna who had come to the meeting to
speak.
And he said, "Don't put living chemists
there, put dead chemists."
And that's what they did; in almost all cases they were
dead.
And it's a wonderful group; as we've already used Faraday
over here.
Okay, here's one of them that's over in that corner of the
building.
And van't Hoff, have you heard of van't Hoff?
You'll have heard a lot of him by the end of this semester.
And Gibbs?
Students: Yes.
Prof: Yes Gibbs, we'll talk about him too.
And Mendeleev you've heard of.
But I suspect that Crookes is not so familiar to you.
This is Sir William Crookes.
He was a Fellow of the Royal Society, and FTS.
In 1861 he discovered the element thallium.
He also developed the cathode ray tube, which he's holding
there, which became the x-ray tube, and he invented the
Cooke's radiometer.
I bet every one of you has seen a Cooke's radiometer.
Do you know?
There's a picture of a Cooke's radiometer, to measure the
intensity of light by how fast it spins.
You've seen these things you buy in novelty stores that sit
there and spin in the light.
Okay?
From 1913 to 1916 he was President of the Royal Society,
the same thing that was founded back by Boyle -- remember?
-- and Robert Hooke and those guys.
In 1898 he was President of the British Association for the
Advancement of Science, and he was also,
in that year, President of the Society for
Psychical Research.
And FTS is Fellow of the Theosophical Society.
In his presidential lecture to the British Association for the
Advancement of Science, one of the things he said was,
"Telepathic research does not yet enlist the interest of
the majority of my scientific brethren."
But he's the guy who supplied the samples for Leadbeater and
Besant to look at.
He supplied them with lithium, chromium,
selenium, titanium, vanadium, boron and beryllium
samples so that they could draw these pictures of what they saw
by clairvoyance.
So you can read more on the Web if you want to about him,
he's an interesting character.
And there were a lot of semi-, well really serious scientists.
Oliver Lodge was another one who invented radio and was head
of the Physics Department at Bristol,
I think , who was into communicating with the dead and
stuff like that.
So it wasn't obvious at the beginning who you should believe
and who you shouldn't believe.
But to get back to the point, and forget clairvoyance for
awhile, which I don't believe in,
the question is, are molecules unobservably
small for what Newton called "vulgar eyes"?
Is there no way we can see something and measure it if it's
that small?
So consider water.
A cubic centimeter of water is 1/18th of a mole,
since the molecular weight is eighteen.
That means it's 6/18^th times 10^23rd molecules,
in a cc of water.
Now that's a really, really big number and it sounds
like they must be just impossibly small,
hopeless to try to see such a thing.
But the difference is between the cube and the cube root,
because when we measure distance we measure linear,
not volume.
So to get something about distances you take the cube root
of the volume.
So if we take the cube root of that, it's about three times
10^7.
Now that's still a very big number, 30,000,000.
Right?
But it's not impossibly small.
It's about -- 3 angstroms is this dimension,
the size of a water molecule.
Remember, we'll talk a lot about a carbon bond --
carbon-carbon bond, being about one and a half
angstroms.
Okay, now think about 10^5.^( )So this lecture room is about
ten meters wide.
One of my hairs is order of magnitude 100 microns in
diameter, and that's a ratio of 10^5.
Right?
So 10^5 of my hairs sideways would go across this room.
So that's a lot.
Right?
But it's not impossible to think of lining up hairs to go
across the room, or putting them side by side.
It'd take a while, but it's not impossible to
think about that, a hair compared to the room.
Right?
But a molecule is about a nanometer, and a small atom is
about 1/10 of a nanometer, or an angstrom.
Right?
And that ratio is 10^5.
So the room to my hair is like my hair to a molecule,
or even an atom, a big atom.
So it's not impossibly small, it's just very,
very small.
In fact, a nucleus is 10 femtometers, which is another
10^5 down.
So that means if the lecture room were an atom,
the nucleus would be the diameter of a hair;
very small, but not impossibly small to conceive of.
Okay?
Now Newton, in Opticks, the book we looked at before,
in 1717 wrote on page 369: "The thickness of the
Plate where it appears very black,
is three eights of the ten hundred thousandth part of an
Inch."
"Is" - that means he knew the size
of it.
Now how big is three-eighths of a ten-hundred-thousandth part of
an inch?
Well ten-hundred-thousand is a million;
three-eights of a millionth of an inch is thirty water
molecules.
So somehow Newton was able to measure something,
in 1717, that was the size of thirty water molecules.
Now what tool could Newton possibly have had in 1717 that
allowed him to measure something so very small?
What?
Student: Glass ring.
Prof: Can't hear very well.
Student: Glass ring.
Prof: Glass springs?
Student: Ring, like you would --
Prof: That's right.
Okay, so here it is, Newton's Rings,
which really should be called Hooke's Rings,
because he's the guy that in Micrographia published
how they worked.
But remember, they had different light
theories.
So here are two pieces of glass, two disks of glass.
And I'll change the lights so we see light through them.
Now I can't remember which is which, but let me guess that if
I put this one over here, having turned it over -- okay.
Oh I can see, yes you can see.
See those patterns?
You've seen things like that with microscope slides,
colors.
Okay?
So these are two flats against one another, they get very close
to one another.
But if I turn the top one over, you see that it's a little bit
-- let me -- so that's what we just saw.
But notice that if I turn it over there's a little --
there turns out here at the edge to be a little bit of a
gap, because the top one is just
very slightly curved.
And now let me see if I can see something.
I can't see it yet, let me zoom in.
We're going to need to focus on that.
Ah, see that thing, see that.
<<Technical adjustment>>
There.
Okay, so here's what you're seeing.
In the middle is what's called Newton's Rings.
They look like that.
Different colors, right?
And Newton could measure the thickness of the air gap that
caused different colors.
The colors repeat in higher orders.
Okay?
Now how could he know the distance?
So he could associate every color with a distance.
How could he know the distance?
Because he could measure the diameter of the rings.
And he said this: "Observation Six:
The Diameter of the fixth Ring" (Right?
One, two, three, four, five, six) "at the
most lucid part of its Orbit was 58/100ths of an inch,
and the Diameter of the Sphere on which the double convex
Object-glass was ground was about 102 Feet;
hence I gathered the thickness of the Air or Aereal Interval of
the Glasses in that Ring."
So in other words, here's the air gap he's trying
to measure, which is the sixth ring out.
He knows the diameter of that ring.
He knows that they touch in the middle, the two glasses,
and he knows that it's a sphere and it has a radius of fifty-one
feet.
So he can just use trigonometry to figure out that the air gap,
if he knows that angle there, the air gap is 1.8 microns,
right?
So that's a lot bigger than an atom;
but if you go into the first ring, or even closer,
then you can measure it.
Right?
So all he needed was a spherical glass with a very
large radius that he could measure his distances from.
Okay, but here's a simpler measure of an even smaller
distance, about 100 years later, and here's the guy who did it.
You know who that is?
<<Students speak over one another>>
Prof: Benjamin Franklin of course;
we'll get back to the artist later on.
So he published in the Philosophical Transactions of
the Royal Society in 1774 -- the Society is now
110-years-old, 114-years-old --
this article on the stilling of waves by means of oil.
Can you see what that means?
What's that related to, a common saying?
Student: Separation?
Prof: Pouring oil on troubled waters;
have you heard of that one?
So the stilling of waves by means of oil.
Extracted from sundry letters between Franklin and Brownrigg.
So he wrote to Brownrigg, 1773: "I had,
when a youth, read and smiled at Pliny's
account of a practice among the seamen of his time,
to still the waves in a storm by pouring oil into the sea;
as well as the use made of oil by the divers…."
Pearl divers would fill their mouth with oil and when they
went down, if a breeze came up and ruffled
the surface, which made it hard to see
because the light wouldn't come through clearly,
then they'd let oil out of their mouths that would float to
the surface, stop the ripples,
and they could see to get their pearls.
This is what Pliny said.
Right?
But he said he smiled at that, what a quaint thing to think;
it's like the occult chemists.
"I think that it has been of late too much the mode to
slight the learning of the ancients.
The learned, too, are apt to slight too much
the knowledge of the vulgar.
In 1757, being at sea in a fleet of ninety-six sail bound
against Louisbourg," (off Cape Breton Island)
"I observed the wakes of two of the ships to be
remarkably smooth, while all the others were
ruffled by the wind, which blew fresh.
Being puzzled with the differing appearance,
I at last pointed it out to our captain and asked him the
meaning of it.
'The cooks,' said he, 'have I suppose just been
emptying their greasy water through the scuppers,
which has greased the sides of those ships a little.'"
(So there's experimental evidence for what Pliny said.)
"Recollecting what I had formerly read in Pliny,
I resolved to make some experiments of the effect of oil
on water when I should have the opportunity."
And when he had the opportunity was when he was in London.
So here's London, and his experiment in 1762 was
in that place, on Clapham Common in South
London.
"At length being at Clapham, where there is on the
common a large pond which I observed one day to be very
rough with the wind, I fetched out a cruet of oil
and dropped a little of it on the water.
I saw it spread itself with surprising swiftness upon the
surface; but the effect of smoothing the
waves was not produced."
(So the experiment was a failure.)
"For I had applied it first on the leeward side of the
pond where the waves were greatest;
and the wind drove my oil back along the shore."
So it didn't spread on the surface.
So what did he do?
Students: Go to the other side.
Prof: Go the other side, right?
"I then went to the windward side where the waves
began to form and there the oil, though not more than a teaspoon
full, produced an instant calm over a
space several yards square which spread amazingly and extended
itself gradually till it reached the lee side,
making all that quarter of the pond,
perhaps half an acre, as smooth as a looking
glass."
So a teaspoon is five cubic centimeters,
and he stilled half an acre, which is 2000 square meters,
which means the thickness of this layer,
to spread that many cubic centimeters over 2000 square
meters, would be twenty-five angstroms,
which is the length of a molecule of the fat that's in
the oil.
So Franklin had in fact measured the length of an oil
molecule, in this way.
Now he didn't think he had measured it and didn't claim to
have measured it.
He said, "When put on the water it spreads instantly many
feet around, becoming so thin as to produce
the prismatic colours" (the Newton Ring's colors,
right?)
"for a considerable space, and beyond them so much
thinner" (it became so thin that you
didn't even get Newton's colors, right?)
"so much thinner as to be invisible except in its
effect of smoothing the waves at much greater distance."
(So the technique was stilling the water;
allowed you to measure how big the area was.
Isn't that cool?)
"It seems as if a mutual repulsion between the particles
took place as soon as it touched the water."
That is, they pressed one another apart and moved.
So what he wouldn't have known was that they pushed one another
apart but they stayed in contact with one another,
to become a monomolecular thick layer over the water.
So he didn't claim to have measured it, but indeed he did
measure the size of an oil molecule.
So molecules are very, very, very small,
but they're not inconceivably small and there are ways you can
measure them.
So we have this question, are there electron pairs?
So let's try first feeling, with Scanning Probe Microscopy,
or SPM.
Can we feel individual molecules?
Can we feel individual atoms?
Can we see what bonds are by feeling them?
So Scanning Tunneling Microscopy was invented in 1981
by these guys in Zurich, Gerd Binnig and Heinrich
Rohrer.
And here's the first publication.
They worked at IBM and here's the first publication on the
cover of their journal about a meeting about this new
technique, Scanning Tunneling Microscopy.
And they got the Nobel Prize within five years,
or maybe it was even less; in very short order, 1986.
Now how does that Scanning Tunneling Microscopy,
which is one type of SPM scanning probe --
Scanning Tunneling is one way, and I'll mention that shortly.
But first I'm going to talk about Atomic Force Microscopy.
And for this you need a little chip, and I'll show you.
The chip is in here, in this little bug box.
Ah you can't see it here.
Well there's the chip in the middle, right?
And I'll pass it around so you can look at it.
And if you tilt it so that the light from the ceiling glints
off the gold, and you see it really shining,
you may be able to see at the thin end these little v's;
you can see here at the bottom little v's.
The chip is 1.4 millimeters wide.
It's a gold-coated silicon chip.
And if you look at those with even higher magnification you
see this -- and there's the size of a hair.
You can actually see -- did you see them?
They're not easy to see and maybe you will and maybe you
won't, but they're there.
Okay, so there's the size of a hair.
And in any given experiment you have a choice of five different
cantilevers, as they're called, that you can use.
Here's a higher resolution electron micrograph that shows
the tip of one of these v's, and you see a little bulge
here, at the bottom, and if you zoom in on that you
see this; it's a little pyramid.
And so at the bottom of that pyramid the radius of curvature
of this tip is only about twenty nanometers, so about the size of
twenty molecules.
Right?
It's round on the tip, it's not truly pointed and come
to nothing; it's a little bit rounded.
You can buy ones that are a third that size,
their radius of curvature.
Okay, and then you have this piece of glass.
I said, "Do not touch" when you came in
because that piece of glass costs $2000.
Okay, so there's this piece of glass, and if I hold it up here
there's a little chip on the bottom of the piece of glass.
I know what I'm looking for.
I don't know if you can see it or not.
But anyhow it's held into this piece of stuff here.
And those little tips are pointing up here,
but in fact the thing is used upside down so they point down.
So here's that thing mounted in an instrument,
and you see there's a red light glowing in the glass,
and the reason is it's being irradiated by --
so there's a blown up picture of one of these tips,
pointing down.
And laser light comes and bounces off the shiny gold on
this tip, and gets reflected up to a detector.
Now what would happen if you pushed up on the tip?
It would deviate the reflection, right?
It would go like this.
And you have a very precise position detector for the laser
light up where you see the light glowing.
So you can tell the deviation of that tip, moving up and down,
by a size that's less than an atom.
Isn't that neat?
So now what you do in this is move this back and forth over
the sample and watch the little needle go up and down by the
reflection of the light.
So if you had, if the surface was like this,
they'll click, click, click,
click, click, click, up, up,
up, and just scan across.
And it's sensitive, as I say, to less than one
molecule change in height; although it doesn't feel an
individual molecule because the tip, remember is twenty
nanometers wide.
So the width of the tip is like 100 or so atoms.
Right?
So it's touching a bunch of atoms at any one time.
But if you come to a ledge, unlike a crystal ,
it would be going at a certain height,
touching a bunch, and then it would move up and
touch a bunch more, and you'd be able to tell if it
went up and down, if that distance was only one
molecule; easy.
Here's, in fact, a picture taken by an
undergraduate.
In fact, it's a movie.
So it's AFM traces.
So what you do, it's like a TV where you go
zoom, zoom, zoom, zoom, zoom, zoom,
zoom, zoom, and then you go back to
the top, zoom, zoom, zoom,
zoom, zoom, say how high it is at every point along the scan,
and then color code it to show how high.
So the width of this is about 600 molecules,
right?
But the lines you see, which are ledges that go up and
down; each of those ledges is one
unit cell, 1.7 nanometers, one set of molecules.
So these are -- and that, the thing at the top surface
where it's flat, is absolutely flat,
except for there's little holes there.
And those holes were made by taking this tip and pressing
down and scratching a little bit to knock some of the molecules
away.
So the larger pit is five nanometers deep,
which means it's four layers of molecules that've been knocked
off in there.
And now this is underwater, or underwater in propanol.
So the crystal is dissolving slowly.
So those pits open up, as molecules come away.
So watch.
Right?
Those are at one-minute intervals.
So we're actually feeling individual molecular heights
with this thing.
There's an interesting point here that I'm not going to dwell
on, but the smallest one did not lose any molecules.
If I back up, watch the smallest one.
It always stays the same size as the others dissolve.
That's an interesting puzzle isn't it?
But at any rate, you can do it.
And you can measure the rate at which different ledges dissolve.
Some dissolve much faster than others.
Here's scanning tunneling microscopy.
This is done in a different way.
It's not -- you don't reflect the laser light.
What you do is detect the electrical conductivity through
a layer.
And if you have an atom there that is a good conductor of
electricity, you get more current through and less and so
on.
So as you scan a real -- now this could be a really,
really sharp tip where the tip is just one atom or a couple of
atoms.
So now as you go across you can feel things that are much
smaller laterally.
Okay?
So this picture here, that thing right there,
which is as you see a repeated pattern,
that's one of those molecules, with 12 carbons and a bromine
on the end.
And here's a model that shows what the molecule looks like.
The yellow is bromine, the brown are the two oxygens,
and the green are carbon and the grey are hydrogens.
They don't come colored that way, they come looking like
this, as to conductivity.
But you can see individual atoms.
Right?
Or one of the more dramatic early examples,
in 1993, was at the IBM labs at Almaden in California,
where they worked on iron atoms on a surface of copper,
and they used this tip and changing the voltage to pick up
atoms and move them where they wanted,
then change the voltage and drop them there,
then go back and get another one and do it.
So they made this thing they called a quantum corral of
whatever, fifty-eight or something iron atoms.
Yes?
Student: Is that along the lines of when they moved the
atoms -- Prof: Pardon me?
Student: Is this along the lines of when they moved the
atoms to spell out IBM?
Prof: Oh yes, they also spelled out IBM.
They knew which side their bread was buttered on too,
just like the Royal Society did.
Okay, so that's neat.
So clearly you can see individual, or feel individual
atoms, but you don't see the bonds.
The bonds are smaller.
It just looks like a bunch of balls, right?
You don't see the electron pairs between them.
Here's SNOM.
SNOM stands for Scanning Near-Field Optical Microscope.
So this is in a sense seeing, but actually it's scanning.
It's like feeling.
So what you have here is a glass fiber that's drawn down to
a very sharp point and coated with aluminum,
and the hole at the bottom is just 100 nanometers;
that's like 100 molecules, not really, really tiny like
the scanning tunneling microscope.
So what you do is you have something on the surface of a
slide here that you want to probe with this,
and you send light down the thing, and if there is a
molecule where the light's coming out,
through this tiny hole, if there's a molecule there,
that will take the blue light and emit red light.
Then you know when it's under the tip.
Suppose there was just one molecule on the whole slide.
You scan it around until suddenly red light comes out.
Then you know that tip is pointed toward that particular
molecule.
Okay?
So red light comes out and you have a detector that detects it,
and you scan the sample back and forth,
like a TV, and you find out where such molecules are.
And here's a picture taken in that way.
So this is a scale of microns, 2 x 2 microns;
in fact it's a distorted scale, as I see here.
Okay?
And that arrow, doubled-headed arrow,
is the wavelength of red light.
So you're seeing things significantly smaller than the
wavelength of the light, having done it in this way.
You're not really looking with your eye, you're doing this
scanning trick.
Okay, so scanning probe microscopies,
like atomic force microscopy, scanning tunneling microscopy,
scanning near-field optical microscopy,
are really powerful, and you can see individual
atoms.
Right?
The sharp points can resolve individual molecules,
and even atoms, but not bonds.
So it's not doing what we need to do for our purposes.
Now, that's -- I'm going to spend the last few minutes going
over the problem that we were looking at with --
remember you were supposed to draw all the resonance
structures for H, C, O, N isomers.
I think you remember that.
Okay, so let's just look at it.
So what we're going to do is compare what we see with Lewis
theory with things that have been calculated by reasonably
good quantum mechanics; not the very,
very highest level, but a pretty good level.
Okay, so we want to try to make all the possible arrangements
here.
So we can put the three atoms, O, C, N, with any one of them
in the middle, and then we can put hydrogen on
either end.
So then we'll do it with nitrogen in the middle and then
with oxygen in the middle, so we'll have done all
possibilities there.
So here are two possibilities.
Now are these good structures is what we have to decide?
Okay, now it would be possible to shift electrons here and draw
a different Lewis structure.
Which of those two do you think is better?
Student: Probably the top one.
Prof: Why?
<<Students speak over one another>>
Prof: Okay, there's separation of charge on
the bottom.
Okay?
So there's bad charge separation.
The most electronegative atom, oxygen, in fact has a positive
charge not a negative charge.
Right?
So that's -- not only is there a charge separation,
it's in the wrong direction.
But they're all octets that we've drawn in both structures.
All atoms have octets.
Now let's try the same trick over here.
We can do it that way.
How about that?
Which of those is better?
In both structures we could go through and count and see that
they're all octets.
Would you like me to do that, or is that something that's
become second nature to you by now?
Anybody want me to count them?
Okay, you've got that.
But now which of those two structures do you think is
better?
<<Students speak over one another>>
Prof: So you say the top one's better,
you say, because it doesn't have charge separation.
Now, as compared to the case on the left, which do you think --
if you have to have charge separation, which one is better?
Student: The right.
Prof: The one on the right because at least you have
it with the right atom, the oxygen being negative.
You could've drawn the one on the top too, right?
And that goes the other way.
Notice that I've introduced something here that I haven't
really talked about enough, the idea of curved arrows.
Curved arrows show a shift of electrons, an electron pair.
What would you draw if you wanted to show just one electron
shifting, if you had to think up the notation?
Students: A dotted curve.
Prof: A dotted curve would be one possibility.
Another possibility?
Sophie?
Student: Half a head instead of a full head.
Prof: Ah, like a fishhook instead of the
double arrow.
And that's in fact what's used.
If you want to shift just one electron you show a single barb,
and if you do a pair of electrons you show a double
barb, a curved arrow.
So curved arrows don't show atoms moving.
They don't show this atom goes to there.
They show an electron pair moving, or sometimes a single
electron moving, and you can use a double or a
single barb to show which one.
Okay, so get that straight, because that's often a problem.
Okay, so the one on top is much worse than the one on the
bottom.
Okay, so there are our possibilities with C in the
middle.
Now how about if you put N in the middle?
Here are two structures with octets.
Now -- oh pardon me, they don't all have octets,
that one on the right has a sestet, because the carbon's
making two bonds; so it gets three electrons out
of those two bonds -- pardon me, it gets two
electrons out of those two bonds,
one from each bond, that's its share from the
bookkeeping, plus an unshared pair.
Right?
That's only -- what's associated with it,
it has only three… pardon me, I'm saying the wrong
thing here.
It has four electrons that it calls its own,
from a bookkeeping point of view, one from each bond and two
in the unshared pair.
So it's not charged.
It came in with four electrons, it's still got four electrons,
but participates with only three pairs of electrons;
so it's a sestet.
So that's not so good on the right.
The one on the left is not so good because it's
charge-separated.
At least it's on oxygen, the negative charge.
We could draw this other one, which puts the negative charge
on carbon.
That's, we would say, not so good.
Right?
So that's a bad charge.
Or we could do it that way.
That looks like the best charge probably;
better to put positive on carbon than on nitrogen.
Okay, but that carbon has a sestet;
the others are octet ones.
Okay, we can do this other one over here.
Now we've got an octet on the carbon, but we've got negative
charge on the carbon.
That's not such a great charge separation thing.
Or we can do it this way, but again we got bad charge.
Okay, so now if we put O in the middle we can do these
structures, both of which you notice have -- what's bad about
them?
<<Students speak over one another>>
Prof: Sestet there, sestet on the right too.
There's a plus charge on oxygen; that's not so great.
You can draw curved arrows and shift the electron pairs around.
You've now got an octet on carbon but you got a sestet on
nitrogen, and bad charge, negative on carbon;
not so great.
And over here you could do that.
Sestet again, bad charge.
Or you can make the three heavy atoms into a ring,
and then you could put the hydrogen on any one of the --
coming out of any one of them.
Now the one on the left, with my computer,
I can't calculate it.
I can't find an energy minimum that has that structure.
So that one is very bad; which doesn't surprise anyone.
It's got horrid angles for the bonds, although we haven't
really said that we prefer one angle over another.
But it also has bad charge separation.
Okay?
That one's a bad charge.
So we could draw this other one here, but it's still got a
positive on oxygen, a sestet.
What do we have here?
A sestet on carbon.
Okay, so here are the bunch of isomers we've just gone through.
Now a couple of years ago, when I first talked about this
in the class, and when these things had been
accurately calc… or as accurately as people are
likely to do, to calculate their true energy
and structure -- I tested my colleagues in the
department, asked them if they could rank
these things as to what's the lowest and what's the highest
energy and so on.
And no one succeeded; in fact, only a few people get
the -- know which the lowest one is, on the basis of their
familiarity with Lewis theory.
So the point is, don't be depressed when you are
trying to do this.
It takes a lot of lore, and even with all of the lore
that's been gathered over many years,
people can't use Lewis theory for this purpose because it's
just not that good.
Okay, but in fact the way I've drawn them there is their
energy, according to these pretty good calculations.
The very lowest energy is the C in the middle and H on the
nitrogen, and then C in the middle and H on the oxygen.
Now you can go back and look at what we said was good and bad
about these things and convince yourself that "Ah!
now you understand why those are on the bottom and these are on
the top."
But if it had come out differently you would have
convinced yourself differently.
Okay?
So that's enough for today.
Actually, there's a reference for this if you want to see it.
The energies were published in the Journal of Chemical
Physics in 2004.
Mobile Analytics