Placeholder Image

Subtitles section Play video

  • The following content is provided under a Creative

  • Commons license.

  • Your support will help MIT OpenCourseWare

  • continue to offer high-quality educational resources for free.

  • To make a donation, or to view additional materials

  • from hundreds of MIT courses, visit MIT OpenCourseWare

  • at ocw.mit.edu.

  • PROFESSOR: So I got much more than one request

  • to do some stuff on nuclear materials,

  • and I think it's just about the right time.

  • That you guys know enough about radiation

  • interacting with matter and everything,

  • and stopping power, and processing,

  • to actually make sense of nuclear materials and radiation

  • damage.

  • And this is my whole theme, so happy

  • to come talk to you guys about this

  • and show you why I think it's interesting.

  • Because it all goes--

  • this slide kind of gets onto it.

  • It starts off with the single-atom atomic defects

  • that make up the basic building blocks of damage

  • and ends up with things that break in nuclear reactors

  • under radiation.

  • And so to understand the whole thing,

  • you've got to know everything from the single atoms

  • on the sort of femtosecond scale,

  • all the way up to the engineering scale

  • where things evolve over years or even decades.

  • So we'll be talking--

  • first, probably today, we're going

  • to go over a material science primer.

  • So who here has had any courses in material science?

  • No one.

  • That's good because I'm assuming that there is a--

  • see, no one knows anything here.

  • I know there's a couple material scientists in the class,

  • and I'll apologize ahead of time if it's a bit of a review.

  • But we'll be going mostly through what are materials

  • and what are the defects that change their material

  • properties, and how do they behave.

  • That'll take us through about today.

  • So then tomorrow, we can see how radiation causes those defects

  • and actually changes material properties.

  • So there's a whole laundry list of different ways

  • that materials fail, and most folks

  • are concerned with all of these--

  • everything from simple overload, which means you

  • stress something too much and it just breaks,

  • to all the different forms of corrosion.

  • That's a whole field in itself.

  • And then there's the things that just

  • we have to worry about because they're only activated

  • with radiation damage.

  • And in this case, this isn't quite ionization by radiation,

  • but it's actual radiation slamming into nuclei

  • and moving atoms out of their place.

  • And we've got one figure that we had recently

  • in a paper that sums up the entire multi-scale picture

  • of radiation damage, from the femtosecond to, let's say,

  • the megasecond scale.

  • Or I think it's more than that.

  • Maybe gigasecond would be the right word for that.

  • And all the way down from the angstrom to the meter scale.

  • And I want to walk you through sort of a lens scale

  • by lens scale depiction of radiation damage.

  • It all starts with knocking atoms out of place.

  • We've mentioned this a little bit

  • when we talked about nuclear stopping power,

  • and this is where it actually comes into play.

  • Sometimes an incoming neutron or photon or ion

  • can displace an atom from its original site,

  • and we call that a physical-- it's a displacement.

  • And then that atom comes off with quite a bit

  • of kinetic energy and can knock into a whole bunch

  • of other atoms.

  • Now this loss of the solid crystalline structure,

  • you can't really tell what the original structure looked like,

  • right?

  • It actually comprises a very small, localized zone

  • of melting called a thermal spike.

  • If you think about, all these atoms

  • are vibrating at fractions of an eV--

  • at thermal energies, like the thermal neutrons

  • we talked about in the reactor.

  • Then you hit them with an MeV neutron.

  • They might transfer 100 keV of energy.

  • And a bunch of these atoms will then

  • be moving about at, let's say, a few hundred eV.

  • That's way beyond liquid temperature.

  • So actually, it's been theorized that there's

  • a little pocket of atoms around three to five nanometers wide

  • that reaches, like, 10,000 Kelvin for a very, very

  • short amount of time--

  • less than a picosecond.

  • Because almost instantly, those atoms

  • knock into the ones around them, and this

  • is how the process of heat transfer occurs.

  • And so, very quickly, you get what's

  • called the quench, where most of those atoms

  • very quickly knock into other ones,

  • slowing down, finding their equilibrium positions again,

  • but not every one.

  • You can see there's a few places where the atoms are still

  • out of their original location.

  • And it's those residual defects that actually

  • comprise radiation damage.

  • And as those defects build up, they start to move.

  • They can diffuse.

  • They can be transported ballistically

  • by more radiation damage.

  • They can move by all sorts of different mechanisms

  • and eventually find each other, forming what's called clusters.

  • So a bunch of those missing atoms could find each other

  • and make a hole, which we call a void.

  • A bunch of the extra atoms shoved in

  • between the other ones can form things

  • called interstitial clusters.

  • We say interstitial because it's like in the space

  • in between where you'd normally find some atoms.

  • So let's say you had a whole bunch of those missing atoms

  • come together, forming a void.

  • This is an actual Transmission Electron Microscope,

  • or TEM, image of a void--

  • pockets of vacuum in materials.

  • Notice anything interesting about its shape?

  • AUDIENCE: It's, like, rounded.

  • PROFESSOR: It's rounded, but what's most striking to me

  • is it isn't actually round.

  • So you would expect a void or a bubble to be kind of spherical,

  • right?

  • That's the minimum energy configuration of most things.

  • Not so when you have a little pocket of vacuum.

  • It's where crystallinity comes into play.

  • And these voids can end up forming superstructures.

  • What curious thing do you notice here?

  • For this whole ensemble of voids.

  • Yeah?

  • AUDIENCE: It seems like they're all in line.

  • PROFESSOR: They are all in the same direction.

  • Kind of funny.

  • That's definitely not an accident, right?

  • That's not like they're randomly aligned.

  • There's a reason for this, that we'll

  • go into in a couple slides.

  • Yeah?

  • AUDIENCE: What's the size scale here?

  • PROFESSOR: The size scale?

  • I think these are on the order of 20 nanometers or so.

  • Yeah, I cropped these images just to get points across.

  • Let's see if it says in the older one.

  • Not quite.

  • Yeah, but these voids can get upwards of tens of nanometers.

  • As small as single atoms.

  • Yeah?

  • AUDIENCE: Sorry, what is this?

  • PROFESSOR: This is the accumulation

  • of radiation defects into what's called voids.

  • Yeah.

  • Don't worry, we'll go over it in more detail again.

  • And if you get little pockets of vacuum in your material,

  • you're not creating or destroying mass.

  • You're just moving it.

  • So those voids, where that mass was has to go somewhere else,

  • and you actually get things that swell

  • in the reactor on their own.

  • They don't change mass but they change volume.

  • They just kind of puff up like Swiss cheese, sometimes

  • upwards of 20% or 30% changes in diameter and length

  • for some tubing.

  • Now if you're depending on these fuel rods

  • being a certain space apart in a reactor

  • and they start to swell, squeezing out the coolant,

  • you lose the ability to cool the reactor.

  • Because then how can you get water around something

  • where the tubes have then swelled together?

  • There's lots of other bad things that can happen,

  • which we'll get into.

  • And so then that's the origin of void swelling.

  • From single missing atoms called vacancies,

  • they can cluster into voids which

  • then cause physical dimensional changes of materials

  • on the scale of centimeters to meters.

  • And that's why we say it's this full multi-scale picture

  • of radiation damage.

  • But to understand, what is damage,

  • you have to know what is an undamaged structure

  • to begin with.

  • So it doesn't make sense to say, how does a structure change,

  • if you don't know how it behaves.

  • So I want to give a very quick primer to material science.

  • And apologies to any material scientists in the room

  • because this is going to seem really basic,

  • but this is a very quick intro to this whole field.

  • I want to go over quickly, what is a crystalline solid?

  • A perfectly undamaged material would

  • be a set of atoms lined up in a very regular lattice

  • and of regular array, where you move over a certain distance

  • and you find another atom.

  • And this extends forever and ever and ever,

  • all the way out to when you reach the free surface.

  • And so this is what we would call an undamaged material.

  • A pristine, perfect, single crystal.

  • By crystal, I mean an arrangement

  • of atoms in a certain direction.

  • So notice here, all of the atoms are lined up in, let's say,

  • some cubic xyz way.

  • That's what we would call one crystal or one grain.

  • You'll hear both of those.

  • And you'll notice also that the arrangement of the atoms

  • tends to determine what the physical objects look like.

  • Or we like to say that form follows

  • structure in material science.

  • So for materials like pyrite, which

  • follows a simple cubic structure, that's the crystals

  • you pull out of the ground.

  • They mimic their atomic configurations

  • in physical centimeter-sized space.

  • For gold atoms, they adopt a slightly different structure.

  • It's still cubic but there is atoms

  • shoved into the cube faces.

  • It's what we call Face-Centered Cubic, or FCC.

  • And you start to see cube-looking structures

  • all over single crystals of gold.

  • Another one, gypsum.

  • It's got a very different type of structure called monoclinic,

  • where none of the sides of this parallelogram are the same

  • and there are some funny angles.

  • But if you look at the arrangement of the atoms

  • and the actual crystals of gypsum that grow,

  • you see a striking similarity, which I find pretty neat.

  • I also want to mention, what is the absence of structure

  • in material science?

  • We call that something that's amorphous.

  • Amorphous means without form.

  • So for example, crystalline indium phosphide

  • would have this regular structure like this.

  • You move over a certain distance,

  • you see another green atom, and so on and so on and so on.

  • In an amorphous material, it can still be a solid,

  • but there is no fixed distance between any certain types

  • of atoms.

  • And radiation can cause a lot of this amorphization

  • by knocking the atoms about and having them freeze

  • in random configurations.

  • This is one of the ways that radiation damage can

  • embrittle materials because--

  • well, we'll get into that.

  • So now let's talk about the defects

  • that can be created in a perfect crystal.

  • The simplest ones, we call point defects.

  • They're zero-dimensional because they're just

  • single atoms out of place.

  • You can have what's called a vacancy, where if you had,

  • let's say, a face-centered cubic lattice of atoms,

  • where you have atoms on every cube corner and every face,

  • if you just pull one out somewhere,

  • we refer to that as a vacancy.

  • A missing atom.

  • It had to go somewhere, though, and we'll

  • get to where it is in just a second.

  • So it might be kind of hard to conceptualize, how do we

  • know that there are missing atoms in all these little cubes

  • or lattices?

  • We do have direct evidence.

  • They're what's called quenching studies,

  • where you can measure the resistance or resistivity

  • of a piece of material after heating it

  • to a certain temperature.

  • Because it turns out that the hotter you make something,

  • the more of those vacancies just naturally occur.

  • You won't actually ever find an absolutely perfect

  • single crystal anywhere in nature,

  • unless you go to zero Kelvin for infinite time,

  • then the atoms arrange themselves thusly.

  • There's always some amount of atomic vibration going on.

  • And there's actually some thermodynamic energy gain

  • to having a few defects in your structure.

  • And that number of defects increases

  • with increasing temperature.

  • Once you get to the melting point of a material,

  • or like right before something melts, you can have up to 1

  • in 10,000 atoms just missing.

  • Moved somewhere else.

  • We call that the thermal equilibrium

  • vacancy concentration.

  • And we can measure that using these resistivity measurements,

  • where you heat materials up to higher and higher temperatures,

  • cool them down suddenly in, like, liquid nitrogen

  • or liquid helium, and measure the change in resistivity.

  • The more defects there are, the harder it is for electrons

  • to flow through.

  • And the only thing that could really be responsible there

  • in a single element would be vacancies.

  • So we do know that these really exist.

  • They can also cluster up.

  • It turns out that every time you have

  • a vacancy in a material, the other atoms

  • move in a little bit towards it, relaxing the pressure they

  • feel from the atoms nearby.

  • And one way for a whole bunch of vacancies

  • to lower the stress of the whole atomic configuration

  • is to cluster together.

  • So if you have a whole bunch of vacancies,

  • they may not allow as much stress accommodation

  • as if they were separate, when they're together.

  • Now you might ask, what happened to the original atoms?

  • You can't just take atoms away and then go nowhere

  • because you can't just destroy matter, right?

  • Unless you turn it into energy, which

  • is what we do in nuclear engineering.

  • So in the material science world,

  • they end up as what's called interstitials,

  • where you kind of have a vacancy created from somewhere

  • that knocks that atom out, and it

  • gets stuck in the next biggest space between some other atoms.

  • And we refer to those as interstitials.

  • And those can cluster up, too, to reduce their total stress

  • in the lattice.

  • They can cluster up into what's called split dumbbell

  • interstitials.

  • Instead of having one extra atom shoved in here,

  • you might rearrange a couple so there's

  • two atoms in the center of a cube instead of one.

  • And that tends to be a lower energy or a more stable

  • configuration.

  • So let's look a little bit at the energetics of these point

  • defects because understanding how they move

  • and why will tell us a lot about how radiation damage happens.

  • So it turns out that interstitials

  • are very hard to make.

  • It's really hard to shove an atom

  • where it doesn't want to be.

  • But once you get it there, it moves very easily.

  • Let's draw a quick, simple cubic lattice

  • to do a little thought experiment

  • and explore why that might be.

  • Let's say I want to shove an interstitial atom in here

  • between these other atoms.

  • Well their electron clouds are going to repel,

  • and it's going to push all the nearby atoms away

  • by just a little bit.

  • And these ones might push the other atoms away

  • by just a little bit, stretching out the lattice,

  • or adding some compressive stress

  • wherever that interstitial is.

  • But then how would it move?

  • What's the biggest barrier it has

  • to overcome to get to the next adjacent location?

  • Well, which direction would it go?

  • Would it go this way?

  • Probably not.

  • There's an atom in the way.

  • So it's going to find the path of least resistance

  • to try to get over here, because like we've talked about

  • before, all atoms are always in motion.

  • Vibrating.

  • Some of them will be energetic enough

  • to squeeze through these two atoms

  • and get over to the next site.

  • And that turns out to be a pretty easy process.

  • We can look at the energy required for an interstitial

  • to move.

  • We notice it's really small fractions of an electron volt,

  • whereas creating them takes two or three electron volts.

  • In atomic land, that's a very high energy penalty.

  • Now let's look at vacancies.

  • They're quite the opposite.

  • They're rather easy to make but they're

  • very hard to move, compared to interstitials.

  • Notice that the energy of movement

  • is about the same as the energy of formation for vacancies.

  • To take an atom out or to pluck it out,

  • you have to break every bond between nearby atoms.

  • So you actually have to put energy in to break those bonds

  • and then remove the atoms somewhere else.

  • Now these things are usually made in pairs,

  • so if you think about how much energy would it

  • take to cause a single radiation damage event where you have

  • one vacancy, which let's say would have been right here,

  • and one interstitial, it takes the sum of these two energies--

  • usually about four electron volts.

  • That's not something that tends to happen

  • chemically or from stress or from something like that.

  • But radiation coming in with hundreds of keV

  • or even MeV neutrons, anything's on the table

  • because it's high enough energy.

  • Yeah?

  • AUDIENCE: What would take about three or four eV?

  • PROFESSOR: So it would take about three or four

  • eV to make a pair of a vacancy and an interstitial.

  • If you just add these two up.

  • It comes usually to about three or four eV, or electron volts.

  • And that's a very difficult thing

  • to do in sort of chemical world, where reactions might proceed

  • with fractions of an electron volt.

  • But when you have MeV neutrons coming in,

  • they do whatever they want.

  • They'll do whatever they will.

  • So someone actually asked me yesterday,

  • what sort of materials can you put in the way of neutrons

  • to stop them from doing damage?

  • And the answer is, pretty much nothing.

  • Fast neutrons tend to travel about 10 centimeters, even

  • in things like steel or water, and they're

  • going to hit what they're going to hit.

  • There's not much you can do but put more things in the way.

  • And we can only get to a certain density with regular matter.

  • And I think osmium has upwards of, like, 22 grams

  • per cubic centimeter density.

  • That's not enough to stop neutrons,

  • even over a considerable distance.

  • Unless you had, like, liquid neutron star,

  • that you could pack nuclei in at a way higher number density,

  • not much you can do.

  • So moving up in the dimensions, there's

  • another type of defect called a dislocation, where

  • it's actually energetically favorable to slide

  • an extra half-plane of atoms in between two sets in here

  • in the crystal lattice, creating a sort of bulged-out structure

  • like you see right here.

  • And dislocations are one of the most important defects

  • in material science and radiation damage.

  • They're what I like to call the agents of plasticity.

  • If you deform a material enough that it doesn't just

  • spring back, then most likely, you were creating and moving

  • dislocations in the material.

  • If you think about a couple of different ways

  • to cause deformation--

  • let's bring our perfect lattice back

  • without all these extra notations.

  • If you want to slide or shear two planes of atoms across,

  • and they're all bonded to each other,

  • what do you physically have to do?

  • How can you get these atoms to slide across each other?

  • What sort of energy do you have to put into it?

  • Yeah?

  • AUDIENCE: [INAUDIBLE] energy.

  • PROFESSOR: Yep.

  • Because all these atoms are bonded to each other,

  • if you want them to move, you have

  • to break every bond on that plane.

  • That's a lot of atomic bonds to break

  • and it's extremely unlikely that that would happen.

  • In fact, if you broke an entire plane of bonds

  • in some material like this, what would you physically do to it?

  • You'd snap it in half.

  • That would be fracture.

  • So if you broke every bond down this plane,

  • you would then have two pieces of this fuel rod.

  • That's usually a pretty high-energy thing to try to do.

  • So instead, if you shove an extra half

  • plane of atoms in there, and the bonds

  • are kind of funny like so, right at that extra half-plane

  • location, then what you can actually do is break one.

  • Let's say you break this one, form the next one,

  • then break this one and form the next one.

  • And for a few atoms to move over,

  • you only have to break a line of bonds, not a plane.

  • So it's much less energy-intensive to get

  • a dislocation to move than to just break something in half.

  • Now you might ask, well, then why do things actually break?

  • Whether or not things deform or break

  • is a balance between this process, which we call slip,

  • and breaking an entire plane of atoms, which we call fracture.

  • So this one's called slip.

  • The other mode is fracture.

  • We would rather materials to form in systems like reactors

  • by slip, just moving a little bit,

  • then just breaking altogether.

  • Unfortunately, when enough radiation hits materials,

  • you can fracture things in a brutal manner,

  • and we'll see what happens then.

  • There's a couple kinds of dislocations.

  • One of them is called a screw dislocation.

  • So imagine you had a whole bunch of sheets of atoms,

  • and you made a cut halfway through that sheet

  • and then moved every plane up by one position.

  • You then got what's called a screw dislocation--

  • kind of a spiral parking garage of atoms surrounding

  • that core right there.

  • You can also have what's called an edge dislocation, which

  • is like the one I've got here on the board

  • right here, where you just have an extra half plane of atoms

  • shoved in right there.

  • So there's two types, and they move in two different ways.

  • The edge dislocation behaves like you may physically expect.

  • If you kind of push like this on two planes of atoms,

  • it moves in the direction you push it.

  • Screw dislocations are kind of screwy.

  • If you push like this, it moves perpendicular.

  • Not going to get into why, but just remember,

  • screw dislocations are fairly screwy in the way

  • that they behave.

  • Not quite intuitive.

  • But that's OK.

  • We don't have to worry about those.

  • And the way that they actually move,

  • like we showed right here, is by what's called glide, or slip,

  • where dislocations can slide just by one plane of atoms

  • or one atomic position in a mechanism that

  • looks something like this.

  • Where, as that dislocation moves,

  • you only have to break a line of bonds

  • and then reform a line of bonds, which

  • is a much easier process than breaking

  • an entire plane at once.

  • It's like you have to break the square root

  • of the same number of bonds.

  • I'm going to skip ahead through some of that.

  • There's one other mechanism of dislocation movement

  • that's important to us in radiation damage

  • and that's called climb.

  • This is when you start to think about,

  • what happens if you have a dislocation, which we'll

  • give this symbol right here, and you also have a vacancy,

  • let's say created by radiation damage.

  • If that vacancy can move, it's going

  • to find the most stressed-out part of this lattice.

  • Most likely, the vacancy will move here.

  • In other words, the atom will move over there,

  • leaving this vacancy over there.

  • It's kind of funny to think, like, what does it

  • mean that a vacancy moves?

  • Has anyone ever done anything with semiconductors

  • and talked about electron and hole movement?

  • OK, yeah.

  • So what does it really mean for a hole to move, right?

  • A hole's not a thing.

  • A vacancy's also not a thing.

  • It's an absence of an atom.

  • But here, we can say that the vacancy moves

  • in this direction when the corresponding atom moves

  • in the exact opposite direction.

  • And then what you've actually done

  • is moved your dislocation up.

  • Instead of moving in the slip direction,

  • you've now moved it in a perpendicular direction.

  • This is usually not possible without things

  • like radiation damage or very high temperature.

  • And then, to make things even crazier,

  • you can also have what's called loops

  • of dislocation, some videos of which I'll actually

  • get to show you.

  • You can have a dislocation that has part edge character, part

  • screw character.

  • If you look at how the atoms are arranged here,

  • you're looking from sort of the top-down.

  • You can see that there's an extra half

  • plane of these white atoms shoved in in the black ones,

  • and this right here would be a completely edge dislocation.

  • You can have a gradual transition,

  • where about 90 degrees later, it looks like a spiral

  • and that's a screw dislocation.

  • And the net effect of that is when

  • you push in this direction on an edge dislocation,

  • it moves that way.

  • When you push this direction on a screw dislocation,

  • it moves that way.

  • So when you stress out a dislocation loop,

  • it just grows.

  • You're not actually creating or destroying matter,

  • but what you're doing is causing this small loop of extra half

  • plane of atoms to grow further and further until it actually

  • reaches some obstacle or the outside of a crystal.

  • And these dislocations can actually

  • feel the force from each other.

  • If I draw a clean one because I think it'll be easier to see--

  • if I draw a small lattice of atoms

  • here and then a dislocation core right there.

  • So that's our dislocation core.

  • This region of space right here is compressively stressed.

  • There's more atoms in that space than there want to be

  • and so it's kind of crammed in there.

  • While this region right here is in what's

  • called tensile stress.

  • There's almost some space, like right here, where

  • there's too few atoms and they kind of want there to be more.

  • And these dislocations can feel neighboring stress fields.

  • Let's say there was another one right over here

  • that had its own compressive stress field.

  • They'll actually repel each other

  • because you don't want to add even more compressive stress

  • to anywhere in this group of atoms.

  • So they'll actually repel each other to the point

  • where, if you get two dislocations too

  • close to each other, they'll what's called pile-up,

  • or they'll refuse to move a bit.

  • So I want to show you some videos.

  • We can actually see these dislocations.

  • In this one, you see that faint line right there originating

  • from this area?

  • That's actually a dislocation loop under stress

  • and that's actually growing.

  • So what you're seeing here is an image

  • of electrons passing through material

  • and looking at regions of different contrast.

  • So wherever there is more atoms or fewer atoms,

  • it looks darker or lighter, and that can tell you

  • what sort of defects there are.

  • You guys all see that faint line right there?

  • Notice how the loop's just growing.

  • It's not like you're moving a line,

  • but you're literally growing a line out

  • of what looks like nothing.

  • There's another one we call a Frank-Read source.

  • It's a source of dislocation loop.

  • So what you're seeing here, each of these lines

  • is a single dislocation.

  • And then right there, you see that loop suddenly form?

  • Let's show you that one again.

  • I'll point on where to look.

  • By stressing out materials, you can actually

  • create additional dislocation loops, right around here.

  • And there it is.

  • You guys see that one?

  • Yeah.

  • Out of what looks like nothing but is actually

  • just a couple of atomic defects, you

  • can create a dislocation loop and allow more plastic

  • deformation to take place, which I think is awesome.

  • Look at this one.

  • Another dislocation source in germanium.

  • It's a little easier to see, also

  • because it's making this sort of spiral set of dislocations

  • a little slower.

  • So you can track its motion a little easier.

  • Notice how they all kind of line up on certain atomic planes.

  • Yeah?

  • AUDIENCE: Does the topology of these things ever change,

  • or is it always just a slow [INAUDIBLE]

  • PROFESSOR: The topology will change.

  • Let's say, if it hits another obstacle

  • or another dislocation, yeah, they

  • can slam into each other and change topology.

  • AUDIENCE: Breaking too [INAUDIBLE]

  • PROFESSOR: All sorts of things, yeah.

  • That's a subject for a whole other class, I'd say.

  • I want to skip ahead to the pile-up

  • because I think this kind of gets the point across.

  • But actually, we can see direct evidence

  • that dislocations feel each other's stress fields.

  • When you get enough of them lined up, they won't overlap.

  • They actually push each other in a kind of dislocation traffic

  • jam.

  • Because what's happening on the atomic level is,

  • they feel each other's stress fields.

  • There might be a source of dislocations further away,

  • but when they get too close to each other,

  • it literally is a dislocation traffic jam.

  • I mean, if you try and hit the car in front of you,

  • the repulsion of the electrons between your and their bumper

  • will prevent the cars from getting a certain distance

  • closer to each other.

  • Same kind of thing here.

  • Moving onto grain boundaries, a two-dimensional defect.

  • Any time you have a perfect crystal

  • of atoms that meets another perfect crystal

  • at a different orientation, or where the atoms

  • are arranged in a different direction,

  • you end up with a boundary between them

  • that we refer to as a grain boundary.

  • So you can actually see, this is a direct physical image

  • of atoms of two different crystals meaning at the grain

  • boundary.

  • Again, taken in the transmission electron microscope.

  • So for those who didn't know, yes,

  • we can see individual atoms and the defects between them.

  • I definitely didn't know that in high school.

  • They didn't even mention that whatsoever.

  • Did you guys ever see images like this?

  • Anyone?

  • Yes?

  • Raise your hand.

  • Just one, OK.

  • So yeah.

  • It's important for you guys to know

  • that we can have direct evidence for all this blackboard stuff

  • because you can see atoms in the transmission electron

  • microscope and see what happens when the two of them meet.

  • You see this kind of regular structure of empty space

  • where this grain boundary meets, right?

  • You can actually model it as a line of 1-D dislocations,

  • because if you take a line of 1-D lines,

  • you end up with a 2-D boundary, which

  • you can see very clearly here.

  • It's almost like there's an extra half plane right there.

  • Another one there, another one there, and another one there.

  • And we call that a tilt grain boundary.

  • Grain boundaries are nice in that they can accommodate lots

  • of these little zero-dimensional defects,

  • moving to them without getting destroyed.

  • So grain boundaries are one of those ways

  • that radiation damage can be removed.

  • And that's one of the reasons why most small-grain materials

  • are really--

  • nano-grain materials are more resistant to radiation damage

  • than large-grain ones because they

  • act as what's called sinks or destroyers of radiation damage.

  • There's another kind of 2-D defect called a twin, where

  • you can actually get a little chunk of atoms sort of switch

  • orientation.

  • And you can see these very clearly in, again, TEM

  • micrographs, and the evidence actually that the twin actually

  • is a different physical arrangement of atoms,

  • even though you can't see the atoms in this little band

  • right there.

  • Look at the way the dislocations line up.

  • Those dislocations tend to line up

  • in energetically-favorable directions, and in this grain,

  • they're all this way, and in the twin,

  • they're all lined up like that.

  • And then finally, there's the most intuitive defect,

  • inclusions.

  • A 3-D piece of some other material inside what

  • would otherwise be a pure material.

  • This one, I actually pulled out of the rotor that powers

  • the Alcator fusion reactor.

  • I was asked to do some analysis to find out,

  • is the structure of that rotor changing,

  • because General Electric who was insuring this rotor said,

  • we don't want to insure it anymore.

  • Thanks for the premiums, but we're not insuring it anymore.

  • And we said, why?

  • And they said, oh, it's structurally unsound.

  • So we said, oh yeah?

  • We'll be back in a year and we'll talk about it.

  • And we did a lot of this work to find out

  • that, actually, the structure hadn't really changed

  • since 1954 when it was made.

  • But what we did also see is we could pop out

  • little precipitates of manganese sulfide.

  • So there's always sulfur in iron,

  • and sulfur tends to be a bad actor when it

  • comes to material properties.

  • You throw manganese into iron to scoop up

  • that sulfur in the form of these little precipitates

  • or inclusions, which we were able to see perfectly

  • when we did an x-ray map, just like the one we

  • did after the first exam.

  • It's like we were looking at Chris' copper silver alloy,

  • mapping out where is the copper and silver.

  • I made this image the same way, mapping out,

  • where is there iron, manganese and sulfur.

  • That's how you can tell what it is.

  • And so dislocations and defects can actually interact.

  • Let's say this is the interaction

  • of a 1-D defect, a dislocation, with a 3-D defect, a void.

  • If you have a material that's deforming plastically,

  • very smoothly, and isn't going to undergo fracture,

  • you want the dislocations to be able to move.

  • If you put anything in their way, they tend to get stuck.

  • It's not easy for that dislocation

  • to shear through a whole bunch of extra atoms.

  • And in some cases, you can stop that motion and favor

  • fracture over slip.

  • So any time you make slip harder,

  • it means that you're making fracture more likely.

  • I didn't say you're making it easier,

  • but you're making it more likely.

  • And you would prefer for materials

  • to deform a little bit by a slip than just break by fracture.

  • So I think now is a good point to go over a few key material

  • properties.

  • All of these are sometimes used to describe the same thing

  • in colloquial speech.

  • That is wrong.

  • Has anyone here thought that, let's say,

  • stiffness or toughness or strength meant the same thing?

  • No.

  • OK.

  • A few people.

  • It's OK.

  • Because it's used wrong all the time in colloquial speech.

  • These actually refer to different material properties

  • with different units.

  • And we're going to go into a little bit about what they

  • are and then show you a few videos

  • to test your intuition about the differences between them.

  • So first, I want to mention what you're seeing right here.

  • It's called a stress-strain curve.

  • Stress is simple.

  • Stress is just a force divided by an area.

  • And usually, the criterion for will

  • a material deform or will it break

  • is does it reach a certain stress.

  • It doesn't matter just how much force you put on it,

  • but it's like, how much force per atom

  • or how much force per area determines whether bonds

  • are going to break.

  • And so on the y-axis is stress.

  • Let's say the amount of force per area we're putting in.

  • And strain is the amount of deformation.

  • So that's stress.

  • And strain is, let's say, the change in length

  • over the original length of some material in what's

  • called the engineering or simplified notation.

  • And so something that is stiff means

  • you can put a lot of force into it

  • but it won't deform very much.

  • That's kind of the easiest property to understand.

  • Is something that's very stiff will

  • have what's called a high Young's modulus,

  • or a high slope right here.

  • Something that's super stiff, like a ceramic,

  • you could really push on it quite a bit,

  • but you won't get it to deform like you would this metal.

  • So the opposite of stiff, I would call compliant.

  • Not soft.

  • This is one of those tricky things right there.

  • Something that's stiff, you try and flex it and it won't flex.

  • Something that's compliant, you put a little bit of force

  • into it and it undergoes some amount of strain.

  • And that slope right there between the stress

  • and the strain, we call the Young's modulus.

  • We also note that this part right here

  • is what's called the elastic region of deformation.

  • By elastic, we mean reversible, or it snaps right back.

  • So right here, when I bend this bar and it snaps right back,

  • that's called elastic deformation.

  • And it's reversible, because you can bend one way

  • and it snaps right back.

  • If I bent it more, which I don't want

  • to do because this is a nice zirconium fuel cladding rod,

  • you would deform it irreversibly.

  • You'd bend it permanently.

  • And to undergo what's called plastic deformation, when

  • you deviate from the slope, and then a little bit more stress

  • can cause a lot more deformation.

  • Have any of you guys ever tried pulling copper wire

  • apart before?

  • That's something I'd recommend you try, for thin wire

  • so you don't cut your hands.

  • What you may notice is that it's awfully hard to get

  • the copper deforming in the first place,

  • but as soon as it starts to stretch, it gets really easy.

  • So this is something I recommend.

  • Go to the electronics shop or wherever

  • and try it out on some really thin copper wire.

  • If it's thick, you'll slice through your fingers

  • and you don't want to do that.

  • Strength, however, that's a different metric.

  • Whereas stiffness describes the slope here,

  • strength describes the height, or the stress at which you

  • start to plastically deform.

  • They're in different units.

  • Stiffness is in stress over strain,

  • whereas strength is given as a stress.

  • So when you hear things like the yield

  • stress or the ultimate tensile strength,

  • that's referring to how strong something is,

  • which may have nothing to do with how stiff it is.

  • Toughness is another property.

  • Toughness is actually kind of like the area under this curve,

  • because if you do a force and apply it over a distance,

  • that's like putting work into the material

  • and it ends up being a unit of energy.

  • So toughness will tell you how much energy

  • you have to put into something before creating

  • a new free surface, otherwise known as fracture.

  • And ductility is how much can you deform it before it breaks.

  • So it would be like this point right here on the strain axis.

  • So I'll give a little bit more examples

  • of what this is all about.

  • Toughness, again, is actually measured

  • as an energy required to form a free surface,

  • or propagate a crack, let's say.

  • Whereas something that's ductile,

  • it doesn't necessarily mean that it's tough.

  • Like, if you have a piece of chewed chewing gum,

  • you can stretch it quite a lot with very little energy.

  • And then you can say it's extremely ductile but not

  • very strong.

  • A piece of copper wire, you can also

  • stretch it an extremely far distance,

  • but it takes more energy to do so.

  • So that's both ductile and strong.

  • And then if you apply that force over a certain distance,

  • stretching out the wire, you can also

  • reveal some of its toughness and how much energy

  • it takes to stretch that wire before it breaks.

  • Hardness is the last material property I want to mention,

  • which is not any of the ones that I showed

  • on the stress-strain curve.

  • Hardness is the resistance to a little bit

  • of plastic deformation.

  • So assuming that you're already here,

  • how much more energy do you have to put in

  • to get the material to deform plastic?

  • So very different material properties.

  • I'll try and mention all what they are.

  • So if we have a stress-strain curve like so,

  • and it follows the elastic region

  • and then deforms plastically, this point

  • here is what we call the yield strength.

  • Whatever that point on the stress axis is.

  • This point right here, our strain to failure,

  • we can use as a measure of ductility.

  • This slope right here refers to the stiffness.

  • And finally, this energy right here

  • is something like the toughness.

  • And the hardness isn't quite on this plot.

  • So I want to see if you guys intuitively

  • understand this, because the next lecture,

  • I'm going to be throwing around the words like stiffness,

  • toughness, ductility, hardness, compliance,

  • hard, soft, whatever, and I want to make sure

  • that you just at least intuitively understand.

  • There's a few videos you may have seen before.

  • Anyone here watch the hydraulic press channel?

  • There we go.

  • Finally, something that half the class does.

  • We're going to predict what's going

  • to happen in each of these cases based on these material

  • properties.

  • So in this case, this is a pressurized cylinder of CO2.

  • It's made of aluminum, which is a very ductile material.

  • It's also a very tough material.

  • How do you think it will deform when smashed?

  • Anyone ever tried this?

  • Squishing aluminum stuff.

  • What happens?

  • AUDIENCE: You compress it.

  • PROFESSOR: You compress it.

  • And then what happens?

  • AUDIENCE: Fracture?

  • PROFESSOR: Will it fracture?

  • AUDIENCE: After a while.

  • PROFESSOR: After a while, OK.

  • If you put a lot of energy into it, eventually,

  • when you reach this strain to failure, it should fracture.

  • But in your personal hands-on experience,

  • does aluminum tend to fracture when you bend it a little bit?

  • AUDIENCE: No.

  • PROFESSOR: So then what words would you use to describe it?

  • Based on this curve right here.

  • Yep?

  • AUDIENCE: Ductile.

  • PROFESSOR: Ductile.

  • I would say ductile and not brittle

  • because you can bend it quite a bit or stretch it quite a bit

  • before it fractures.

  • How about stiffness?

  • Is it really hard or really easy to get aluminum bending?

  • AUDIENCE: It's pretty easy.

  • PROFESSOR: It's fairly easy.

  • So would you call that stiff or compliant?

  • AUDIENCE: Compliant.

  • PROFESSOR: Compliant.

  • OK.

  • What about strength?

  • How hard is it to start deforming aluminum

  • irreversibly, compared to something like steel?

  • AUDIENCE: Not very.

  • PROFESSOR: Not very.

  • Especially pure aluminum.

  • You can chew through it.

  • If you guys ever got a one yen coin from Japan,

  • you can chew through it.

  • Not very strong.

  • Then again, your bite force is also incredibly strong.

  • But anyway, let's see what actually

  • happens when you compress a rather ductile, compliant,

  • and not that strong aluminum canister.

  • Is it actually going?

  • Oh, it actually skipped ahead.

  • That's what I wanted, was their sound.

  • It was also pressurized with CO2.

  • But notice what's left.

  • So actually watch in slow-mo.

  • Look how much you can compress that, even after the explosion.

  • No fracture.

  • If you had done that with, let's say, a glass canister,

  • what do you guys think would have happened?

  • AUDIENCE: It would have shattered.

  • PROFESSOR: It would have shattered.

  • Yeah, we'll see that in a bit with a material

  • that may surprise you.

  • AUDIENCE: So it basically doesn't fracture, right?

  • PROFESSOR: It will fracture eventually,

  • but the hydraulic press can't get it that far in compression.

  • So that would be something that's extremely ductile, not

  • that strong--

  • so it wasn't that hard to deform.

  • Certainly we know it wasn't stronger than the steel base

  • plate that they used to do the smashing.

  • Because whatever's the softer material

  • is going to deform more.

  • So here he's going to have-- well I'll let him describe it,

  • and then I'll let you guess what's going to happen.

  • What do you guys think is going to happen?

  • You've got what looks like brass and copper coins

  • on a steel base plate.

  • Anyone have any idea?

  • AUDIENCE: [INAUDIBLE]

  • PROFESSOR: Yeah.

  • Everyone's making this motion, which

  • means everything's going to flatten out, right?

  • Let's find out.

  • Not nearly as much as you might have expected.

  • Is anyone surprised by this?

  • What happened there?

  • What actually happened there was already described up here.

  • When you get enough dislocations piling up against each other

  • during plastic deformation, you can undergo a process

  • called work hardening.

  • That process can be physically described

  • by a lot of those dislocations piling up and making

  • it more and more difficult to continue that deformation.

  • So what happened here is the brass and the copper,

  • which started out quite soft, not that hard, quite ductile,

  • as you can see, and not that strong actually got stronger

  • as they were deformed.

  • Interesting, huh?

  • Did anyone expect this to happen?

  • OK.

  • Let's go to one that I think everyone can guess what's

  • going to happen, a lead ball.

  • So has anyone ever tried playing with lead before?

  • Hopefully not.

  • I have quite a--

  • OK good, I'm not alone.

  • How would you describe lead in terms of the material

  • properties here?

  • AUDIENCE: [INAUDIBLE]

  • PROFESSOR: Yep.

  • It's not very stiff.

  • It doesn't take much energy to start deforming it.

  • How else?

  • Was it hard or soft?

  • AUDIENCE: Soft.

  • PROFESSOR: OK.

  • Do you think it's ductile or brittle?

  • Yeah?

  • AUDIENCE: It's brittle.

  • PROFESSOR: You think it's brittle.

  • So by that, you mean it's just going to break apart, right?

  • If you deform it?

  • OK, cool.

  • And would you say it is tough or not tough?

  • Not a lot of folks have hands-on experience with lead.

  • It's probably good for your brains.

  • Let's find out.

  • Lead pancake.

  • So what words would you use to describe what just happened?

  • AUDIENCE: It's ductile.

  • PROFESSOR: Ductile indeed.

  • I don't know what sort of brittle lead--

  • was it an alloy that you had been playing with, maybe?

  • AUDIENCE: It was like a little sheet.

  • It was just easy to snap.

  • PROFESSOR: Aha, OK.

  • So it was a sheet of lead that was easy to snap.

  • So I would not call lead as a very tough material

  • because you didn't have to put a lot of energy into it,

  • but did it deform quite a bit before you snapped it or did

  • it just crumble apart?

  • AUDIENCE: Oh, it deformed.

  • PROFESSOR: OK.

  • So in that case, I would call it ductile because it deformed

  • a lot before breaking, but I would not

  • call it tough because it took very little energy to get it

  • to that breaking point.

  • And it wasn't that stiff because it was quite easy to get it--

  • let's say it's the amount of stress

  • you put in versus the strain.

  • It could be quite low.

  • And it would not be very strong because it

  • didn't take a lot of energy or stress to get it moving.

  • Let's look at another ball.

  • In this case, a steel ball bearing.

  • What do you guys think is going to happen here?

  • AUDIENCE: It's going to shatter.

  • PROFESSOR: It's going to shatter.

  • So you're guessing that the steel is brittle, right?

  • What else?

  • AUDIENCE: Probably pretty stiff and strong.

  • PROFESSOR: Probably quite stiff and strong, yeah.

  • I think so, too, but I don't think the guy

  • that did this expected that.

  • [INTERPOSING VOICES]

  • PROFESSOR: Yeah.

  • Did that surprise anybody?

  • Yeah.

  • Quite a surprise, right?

  • So in this case, materials like hardened steel

  • aren't necessarily that brittle.

  • In fact, you wouldn't want a ball bearing to be brittle.

  • If you get some small chip in it or a little bit of grit or sand

  • in the bearings, you would shatter the ball bearing

  • and cause instantaneous failure of the rotating component.

  • So what you actually want out of a high-strength ball bearing

  • is something that's extremely hard.

  • Resists deformation so it doesn't undergo,

  • let's say, change of shape that would prevent it

  • from rolling without friction or with very little friction.

  • You want it to be quite stiff because you don't

  • want the load of whatever you're loading onto it to deform it,

  • but you also don't want it to be brittle.

  • So it's got to be somewhat tough and ductile to prevent

  • sudden failure.

  • You'd rather it compress a tiny bit than just cracking in half.

  • So you can make things like ceramic ball bearings, which

  • are very brittle, very stiff, not that tough, but also very

  • strong, and you just have to make sure

  • that whatever part you make is not

  • going to reach any sort of yield strength criterion or crack

  • or anything.

  • Now the last one that's probably the most surprising.

  • They bought a $4,000 diamond.

  • It's a diamond like that big.

  • What do you know about diamonds as a material in terms

  • of these properties?

  • AUDIENCE: They're hard.

  • PROFESSOR: Yep, both is right.

  • They're extremely stiff.

  • It's the hardest material that we know of, almost.

  • We've made slightly harder ones artificially.

  • It's the hardest natural material we know of.

  • What else?

  • Do you know whether they're strong or tough?

  • AUDIENCE: They're not tough.

  • PROFESSOR: They're not tough.

  • Why do you say that?

  • AUDIENCE: Because it will shatter.

  • PROFESSOR: Have you seen the video?

  • AUDIENCE: [INAUDIBLE]

  • PROFESSOR: Oh wow.

  • OK.

  • What else do we have?

  • Yeah.

  • So you're saying it's not tough.

  • AUDIENCE: You can cut diamonds, right?

  • PROFESSOR: You can cut diamonds with other diamonds.

  • So the cutting action usually depends

  • on the relative hardness of the material.

  • So if you want to polish or cut something abrasively,

  • you need to use a harder material

  • because then the grit itself won't wear away

  • before the material it's trying to cut.

  • But what's going to happen here is we're going to put a diamond

  • and try compressing it, and we'll

  • see what its stress-strain curve looks like.

  • So votes on what's going to happen.

  • Who says, like Monica, it's going to shatter?

  • Who thinks it's going to break the tools?

  • Who thinks it's going to deform plastically?

  • Yeah, I've never seen a diamond deform plastically.

  • AUDIENCE: They still have pretty big chunks, though.

  • PROFESSOR: Oh yeah, they could probably still sell those.

  • Absolutely no deformation.

  • It just rotates and explodes.

  • Yeah.

  • This would be a material that we would

  • say has almost zero ductility.

  • Despite being extremely hard, I don't

  • know if there would even been enough

  • deformation to have a slight dent in the tool itself.

  • There's probably a little hole where the point of the diamond

  • poked in, but once there was enough stress on that diamond,

  • its stress-strain curve would look something like that.

  • Maybe like that.

  • Yeah.

  • So it's important that you intuitively

  • understand the differences between strength, ductility,

  • hardness, toughness, and stiffness, because then

  • next class, we can explain how radiation changes them.

  • So any questions on the materials and properties

  • from today?

  • Yeah?

  • AUDIENCE: Can you clarify why something is, for example,

  • ductile versus brittle?

  • PROFESSOR: Mhm.

  • So the reason something would be ductile

  • versus brittle is whether or not you can plastically deform it,

  • and that means whether or not it's more energetically

  • favorable for dislocations to keep moving

  • versus just breaking a plane of atoms in any irregular

  • direction and causing fracture.

  • So again, ductility versus embrittlement

  • is the interplay between slip and fracture.

  • Slip is normally done by dislocation movement.

  • Any defects created by anything, especially radiation damage,

  • will make slip harder so that any continued energy you put in

  • will not move dislocations but move towards fracture.

  • If there's no other questions, we'll

  • look at the stress-strain curves of

  • some other familiar materials.

  • It is 10:00, in case you guys have to go to other classes.

  • AUDIENCE: Are you taking any nuclear activation stuff today?

  • PROFESSOR: Yes.

  • If you guys have things for a nuclear activation analysis,

  • hand it in.

  • You guys bring stuff in?

  • We're running out of opportunities to do this.

  • All right.

  • In that case, the entry fee for the quiz

  • will be your nuclear activation analysis sample.

The following content is provided under a Creative

Subtitles and vocabulary

Click the word to look it up Click the word to find further inforamtion about it