Tag Archives: amyloid

4 diseases explained at one blow said the protein chemist — part 1

A brilliant paper [ Science vol. 377 eabn5582 pp. 1 –> 20 ’22 ] explains how changing a single amino acid (proline) to another  can cause 4 different diseases, depending on the particular protein it is found in (and which proline of many is changed).

There is so much in this paper that it will take several posts to go over it all.  The chemistry in the paper is particularly fine.  So it’s back to Biochemistry 101 and the alpha helix and the beta sheet.

Have a look at this

https://cbm.msoe.edu/teachingResources/proteinStructure/secondary.html

If you can tell me how to get a picture like this into a WordPress post please make a comment.

The important point is that hydrogen bonds between the amide hydrogen of one amino acid and the carbonyl group of another hold the alpha helix and the beta pleated sheet together.

Enter proline : p//en.wikipedia.org/wiki/Proline.  Proline when not embedded in a protein has a hydrogen on the nitrogen atom in the ring.  When proline is joined to another amino acid by a peptide bond in a protein, the hydrogen on the nitrogen is no longer present.  So the hydrogen bond helping to hold the two structures (alpha helix and beta sheet) is no longer present at proline, and alpha helices and beta sheets containing proline are not has stable.  Prolines after the fourth amino acid of the alpha helix (e. g. after the first turn of the helix) produce a kink.  The proline can’t adopt the alpha helical configuration of the backbone and it can’t hydrogen bond.

But it’s even worse than that (and this observation may even be original).  Instead of a hydrogen bonding to the free electrons of the oxygen in the carbonyl group you have the two electrons on the nitrogen jammed up against them.  This costs energy and further destabilizes both structures.

Being a 5 membered ring which contains the alpha carbon of the amino acid, proline in proteins isn’t as flexible as other amino acids.

This is why proline is considered to be a helix breaker, and is used all the time in alpha helices spanning cellular membranes to cause kinks, giving them more flexibility.

There is much more to come — liquid liquid phase separation, prion like domains, low complexity sequences, frontotemporal dementia with ALS, TDP43, amyloid, Charcot Marie Tooth disease and Alzheimer’s disease.

So, for the present stare at the link to the diagram above.

Amyloid Structure at Last ! – 2 Birefringence

This was the state of the art 19 years ago in a PNAS paper (vol. 99 pp. 16742 – 16747 ’02).  “Amyloid fibrils are filamentous structures with typical diameters of 10 nanoMeters and lengths up to several microns.  No high resolution molecular structure of an amyloid fibril has yet been determined experimentally because amyloid fibrils are noncrystalline solid materials and are therefore incompatible with Xray crystallography and liquid state NMR.”

Well solid state NMR and cryo electron microscopy have changed all that and we now have structures for many amyloids at near atomic resolution.  It’s probably behind a pay wall but look at Cell vol. 184 pp. 4857 – 4873 ’21 if you have a chance.  I’ve spent the last week or so with it, and a series of posts on various aspects of the paper will be forthcoming.  The paper contains far too much to cram into a single post.

So lacking an Xray machine to do diffraction, what did we have 57 years ago when I started getting seriously interested in neurology?  To find amyloid we threw a dye called Congo Red on a slide, found that it bound amyloid and became birefringent when it did so.

Although the Cell paper doesn’t even mention Congo Red, the structure of amyloid they give explains why this worked.

What is birefringence anyway?  It means that light moving through a material travels at different speeds in different directions.  The refractive index of a material is the relative speed of light through that material versus the speed of light in a vacuum.   Stand in a shallow pool.  Your legs look funny because light travels slower in water than in air (which is nearly a vacuum).

Look at the structure of Congo Red — https://en.wikipedia.org/wiki/Congo_red.  It’s a long thin planar molecule, containing 6 aromatic rings, kept planar with each other by pi electron delocalization.

The previous post contained a more detailed description of amyloid — but suffice it to say that instead of wandering around in 3 dimensional space, the protein backbone in amyloid is confined to a single plane 4.8 Angstroms thick — here’s a link — https://luysii.wordpress.com/2021/10/11/amyloid-structure-at-last/

Plane after plane stacks on top of each other in amyloid.  So a micron (which is 10,000 Angstroms) can contain over 5,000 such planes, and an amyloid fibril can be several microns long.

It isn’t hard to imagine the Congo Red molecule slipping between the sheets, making it’s orientation fixed.  Sounds almost pornographic doesn’t it? This orients the molecule and clearly light moving perpendicular to the long axis of Congo Red will move at a different speed than light going parallel to the long axis of Congo Red, hence its birefringence when the dye binds amyloid.

Well B-DNA (the form we all know and love as the double helix) has its aromatic bases stacked on top of each other every 3.4 Angstroms.  So why isn’t it birefringent with Congo Red?  It has a persistence length of 150 basePairs or about .05 microns, which means that the average orientation is averaged out, unlike the amyloid in a senile plaque

There is tons more to come.  The Cell paper is full of fascinating stuff.

Amyloid structure at last !

As a neurologist, I’ve been extremely interested in amyloid  since I started in the late 60s.  The senile plaque of Alzheimers disease is made of amyloid.  The stuff was insoluble gunk. All we had back in the day was Xray diffraction patterns showing two prominent reflections at 4 and 9 Angstroms, so we knew there was some sort of repetitive structure.

My notes on papers on the subject over the past 20 years contain  about 100,000 characters (but relatively little enlightenment until recently).

A while ago I posted some more homework problems — https://luysii.wordpress.com/2021/09/30/another-homework-assignment/

Homework assignment #1:  design a sequence of 10 amino acids which binds to the same sequence in the reverse order forming a plane 4.8 Angstroms thick.

Homework assignment #2 design a sequence of 60 amino acids which forms a similar plane 4.8 Angstroms thick, such that two 60 amino acid monomers bind to each other.

Feel free to use any computational or theoretical devices currently at our disposal, density functional theory, force fields, rosetta etc. etc.

Answers to follow shortly

Hint:  hundreds to thousands of planes can stack on top of each other.

 

If you have a subscription to Cell take a look at a marvelous review full of great pictures and diagrams [ Cell vol. 184 pp. 4857 – 4873 ’21 ].

 

Despite all that reading I never heard anyone predict that a significantly long polypeptide chain could flatten out into a 4.8 Angstrom thick sheet, essentially living in 2 dimensions.  All the structures we had  (alpha helix, beta pleated sheet < they were curved >, beta barrel, solenoid, Greek key) live in 3 dimensions.

 

 

So amyloid is not a particular protein, but a type of conformation a protein can assume (like the structures mentioned above).

 

 

So start with NH – CO – CHR.  NH  CO and C in the structure all lie in the same plane (the H and the side chain of the amino acid < R >  project out of the plane).

 

Here’s a bit of elaboration for those of you whose organic chemistry is a distant memory.  The carbon in the carbonyl bond (CO) has 3 bonding orbitals in one plane 120 degrees apart, with the 4th orbital perpendicular to the plane — this is called sp2 hybridization.  The nitrogen can also be hybridized to sp2.  This lets the pair of electrons above the plane roam around moving toward the carbon.  Why is this good?  Because any time you let electrons roam around you increase their entropy (S) and anything increasing entropy lowers their free energy (F)which is given by the formula F = H – TS where H is enthalpy (a measure of bond strength, and T is the absolute temperature in Kelvin.

 

 

So N and CO are in one plane, and so are the bonds from  N and C to the adacent atoms (C in both cases).

 

You can fit the plane atoms into a  rectangle 4.8 Angstroms high.  Well that’s one 2 dimensional rectangle, but the peptide bond between NH and CO in adjacent rectangles allows you to tack NH – CO – C s together while keeping them in a 3 dimensional parallelopiped 4.8 Angstroms high.

 

 

Notice that in the rectangle the NH and CO bonds are projecting toward the top and bottom of the rectangle, which means that in each plane  NH – CO – CHR s, the NH and CO are pointing out of the 2 dimensional plane (and in opposite directions to boot). This is unlike protein structure in which the backbone NHs and COs hydrogen bond to each other.  There is nothing in this structure for them to bond to.

 

 

What they do is hydrogen bond to another 3 dimensional parallelopiped (call it a sheet, but keep in mind that this is NOT the beta sheet you know about from the 3 dimensional structures of proteins we’ve had for years).

 

 

So thousands of sheets stacked together form the amyloid fibril.

 

Where does the 9 Angstrom reflection of cross beta come from?  Consider the  [ NH – CHR – CO ]  backbone as it lies in the 4.8  thick plane (I never thought such a thing would be even possible ! ).  It curves around like a snake lying flat.  Where are the side chains?  They are in the 4.8 thick plane, separating parts of the meandering backbone from each other — by an average of 9 Angstroms

 

Here is an excellent picture of the Alzheimer culprit — the aBeta42 peptide as it forms the amyloid of the senile plaque

 

 

You can see the meandering backbone and the side chains keeping the backbone apart.

 

 

That’s just the beginning of the paper, and I’ll have lots more to say about amyloid as I read further.   Once again, biology instructs chemistry and biochemistry giving it more “things in heaven and earth, Horatio, than are dreamt of in your philosophy.”

Proteins (and amyloids) still have some tricks up their sleeves

We all know that amyloids are made of beta sheets stacked on top of each other. Not all of them, says Staph Aureus according to PNAS e2014442118 ’21. In fact one protein they produce (Phenol Soluble Modulin alpha 3 (PSMα3)– PSMalpha3 ) which is toxic to human immune cells forms amyloid made of alpha helices.  PSMalpha3 forms cross-α amyloid fibrils that are composed entirely of amphipathic α-helices. The helices stack perpendicular to the fibril axis into mated “sheets”

However other members of the family namely PSMα1 and PSMα4 adopt the classic amyloid ultrastable cross-β architecture and are likely to serve as a scaffold rendering the biofilm a more resistant barrier.

It gets worse.

Consider an antimicrobial peptide (AMP) called uperin 3.5, secreted on the skin of a frog which also forms amyloid fibrils made of alpha helices.  The amyloid is  essential for uperin 3.5’s  toxic activity against the Gram-positive bacterium Micrococcus luteus.

It gets even worse.  

When secreted onto the frog skin uperin 3.5. has a disordered structure. Uperin 3.5 requires bacterial membranes to form the toxic amyloid made of alpha helices.   When no membranes are around, uperin 3.5. still forms amyloid, but this time the amyloid is of the classic beta sheet.  So one protein can form two types of amyloid.  Go figure

Uperin 3.5 is a classic example of a chameleon protein. 

Amyloid

Amyloid goes way back, and scientific writing about has had various zigs and zags starting with Virchow (1821 – 1902) who named it because he thought it was made out of sugar.  For a long time it was defined by the way it looks under the microscope being birefringent when stained with Congo red (which came out 100 years ago,  long before we knew much about protein structure (Pauling didn’t propose the alpha helix until 1951).

Birefringence itself is interesting.  Light moves at different speeds as it moves through materials — which is why your legs look funny when you stand in shallow water.  This is called the refractive index.   Birefringent materials have two different refractive indexes depending on the orientation (polarization) of the light looking at it.  So when amyloid present in fixed tissue on a slide, you see beautiful colors — for pictures and much more please see — https://onlinelibrary.wiley.com/doi/full/10.1111/iep.12330

So there has been a lot of confusion about what amyloid is and isn’t and even the exemplary Derek Lowe got it wrong in a recent post of his

“It needs to be noted that tau is not amyloid, and the TauRx’s drug has failed in the clinic in an Alzheimer’s trial.”

But Tau fibrils are amyloid, and prions are amyloid and the Lewy body is made of amyloid too, if you subscribe to the current definition of amyloid as something that shows a cross-beta pattern on Xray diffraction — https://www.researchgate.net/figure/Schematic-representation-of-the-cross-b-X-ray-diffraction-pattern-typically-produced-by_fig3_293484229.

Take about 500 dishes and stack them on top of each other and that’s the rough dimension of an amyloid fibril.  Each dish is made of a beta sheet.  Xray diffraction was used to characterize amyloid because no one could dissolve it, and study it by Xray crystallography.

Now that we have cryoEM, we’re learning much more.  I have , gone on and on about how miraculous it is that proteins have one or a few shapes — https://luysii.wordpress.com/2010/08/04/why-should-a-protein-have-just-one-shape-or-any-shape-for-that-matter/

So prion strains and the fact that alpha-synuclein amyloid aggregates produce different clinical disease despite having the same amino acid sequence was no surprise to me.

But it gets better.  The prion strains etc. etc may not be due to different structure but different decorations of the same structure by protein modifications.

The same is true for the different diseases that tau amyloid fibrils produce — never mind that they’ve been called neurofibrillary tangles and not amyloid, they have the same cross-beta structure.

A great paper [ Cell vol. 180 pp. 633 – 644 ’20 ] shows how different the tau protofilament from one disease (corticobasal degeneration) is from another (Alzheimer’s disease).  Figure three shows the side chain as it meanders around forming one ‘dish’ in the model above.  The meander is quite different in corticobasal degeneration (CBD) and Alzheimers.

It’s all the stuff tacked on. Tau is modified on its lysines (some 15% of all amino acids in the beta sheet forming part) by ubiquitination, acetylation and trimethylation, and by phosphorylation on serine.

Figure 3 is worth more of a look because it shows how different the post-translational modifications are of the same amino acid stretch of the tau protein in the Alzheimer’s and CBD.  Why has this not been seen before — because the amyloid was treated with pronase and other enzymes to get better pictures on cryoEM.  Isn’t that amazing.  Someone is probably looking to see if this explains prion strains.

The question arises — is the chain structure in space different because of the modifications, or are the modifications there because the chain structure in space is different.  This could go either way we have 500+ enzymes (protein kinases) putting phosphate on serine and/or threonine, each looking at a particular protein conformation around the two so they don’t phosphorylate everything — ditto for the enzymes that put ubiquitin on proteins.

Fascinating times.  Imagine something as simple as pronase hiding all this beautiful structure.

 

 

Barking up the wrong therapeutic tree in Alzheimer’s disease

Billions have been spent by big pharma (and lost) trying to get rid of the senile plaque of Alzheimer’s disease.  The assumption has always been that the plaque is the smoking gun killing neurons.  But it’s just an assumption which a recent paper has turned on its ear [ Proc. Natl. Acad. Sci. vol. 116 23040 – 23049 ’19 ]

It involves a protein, likely to be a new face even to Alzheimer’s cognoscenti.  The protein is called SERF1A (in man) and MOAG-4 in yeast. It enhances amyloid formation, the major component of the senile plaque.  SERF1A is clearly doing something important as it has changed little from the humble single yeast cell to man.

The major component of the senile plaque is the aBeta peptide of 40 and/or 42 amino acids.  It polymerizes to form the amyloid of the plaque.  The initial step of amyloid formation is the hardest, getting a bunch of Abeta peptides into the right conformation (e.g. the nucleus) so others can latch on to it and form the amyloid fiber.   It is likely that the monomers and oligomers of Abeta are what is killing neurons, not the plaques, otherwise why would natural selection/evolution keep SERF1A around?

So, billions of dollars later, getting rid of the senile plaque turns out to be a bad idea. What we want to do is increase SERF1A activity, to sop up the monomers and oligomers. It is a 180 degree shift in our thinking. That’s the executive summary, now for the fascinating chemistry involved.

First the structure of SERF1A — that is to say its amino acid sequence.  (For the nonChemists — proteins are linear string of amino acids, just as a word is a linear string of characters — the order is quite important — just as united and untied mean two very different things). There are only 68 amino acids in SERF1A of which 14 are lysine 9 are arginine 5 Glutamic acid and 5 Aspartic acid.  That’s interesting in itself, as we have 20 different amino acids, and if they occurred randomly you’d expect about 3 -4 of each.  The mathematicians among you should enjoy figuring out just how improbable this compared to random assortment. So just four amino acids account for 33 of the 68 in SERF1A  Even more interesting is the fact that all 4 are charged at body pH — lysine and arginine are positively charged because their nitrogen picks up protons, and glutamic and aspartic acid are negatively charged  giving up the proton.

This means that positive and negative can bind to each other (something energetically quite favorable).  How many ways are there for the 10 acids to bind to the 23 bases?  Just 23 x 22 x 21 X 20 X 19 X 18 x 17 x 16 x 15 x 14 or roughly 20^10 ways.  This means that SERF1A doesn’t have a single structure, but many of them.  It is basically a disordered protein.

The paper shows exactly this, that several conformations of SERF1 are seen in solution, and that it binds to Abeta forming a ‘fuzzy complex’, in which the number of Abetas and SERF1s are not fixed — e.g. there is no fixed stoichiometry — something chemists are going to have to learn to deal with — see — https://luysii.wordpress.com/2018/12/16/bye-bye-stoichiometry/.  Also different conformations of SERF1A are present in the fuzzy complex, explaining why it has such a peculiar amino acid composition.  Clever no?  Let’s hear it for the blind watchmaker or whatever you want to call it.

The paper shows that SERF1 increases the rate at which Abeta forms the nucleus of the amyloid fiber.  It does not help the fiber grow.  This means that the fiber is good and the monomers and oligomers are bad.  Not only that but SERF1 has exactly the same effect with alpha-synuclein, the main protein of the Lewy body of Parkinsonism.

So the paper represents a huge paradigm shift in our understanding of the cause of at least 2 bad neurological diseases.   Even more importantly, the paper suggests a completely new way to attack them.

RIPK1

The innate immune system is intrinsically fascinating, dealing with invaders long before antibodies or cytotoxic cells are on the scene.  It is even more fascinating to a chemist because it works in part by forming amyloid inside the cell.  And you thought amyloid was bad.

The system becomes even more fascinating because blocking one part of it (RIPK1) may be a way to treat a variety of neurologic diseases (ALS, MS,Alzheimer’s, Parkinsonism) whose treatment could be improved to put it mildly.

One way to deal with an invader which has made it inside the cell, is for the cell to purposely die.  More and more it appears that many forms of cell death are elaborately programmed (like taking down a stage set).

Necroptosis is one such, distinct from the better known and studied apoptosis.   It is programmed and occurs when a cytokine such as tumor necrosis factor binds to its receptor, or when an invader binds to members of the innate immune system (TLR3, TLR4).

The system is insanely complicated.  Here is a taste from a superb review — unfortunately probably behind a paywall — https://www.pnas.org/content/116/20/9714 — PNAS vol. 116 pp. 9714 – 9722 ’19.

“RIPK1 is a multidomain protein comprising an N-terminal kinase domain, an intermediate domain, and a C-terminal death domain (DD). The intermediate domain of RIPK1 contains an RHIM [receptor interacting protein (rip) homotypic interaction motif] domain which is important for interacting with other RHIM-containing proteins such as RIPK3, TRIF, and ZBP1. The C-terminal DD mediates its recruitment by interacting with other DD-containing proteins, such as TNFR1 and FADD, and its homodimerization to promote the activation of the N-terminal kinase domain. In the case of TNF-α signaling, ligand-induced TNFR1 trimerization leads to the assembly of a large receptor-bound signaling complex, termed Complex I, which includes multiple adaptors (TRADD, TRAF2, and RIPK1), and E3 ubiquitin ligases (cIAP1/2, LUBAC complex).”

Got that?  Here’s a bit more

“RIPK1 is regulated by multiple posttranslational modifications, but one of the most critical regulatory mechanisms is via ubiquitination. The E3 ubiquitin ligases cIAP1/2 are recruited into Complex I with the help of TRAF2 to mediate RIPK1 K63 ubiquitination. K63 ubiquitination of RIPK1 by cIAP1/2 promotes the recruitment and activation of TAK1 kinase through the polyubiquitin binding adaptors TAB2/TAB3. K63 ubiquitination also facilitates the recruitment of the LUBAC complex, which in turn performs M1- type ubiquitination of RIPK1 and TNFR1. M1 ubiquitination of Complex I is important for the recruitment of the trimeric IκB kinase complex (IKK) through a polyubuiquitin-binding adaptor subunit IKKγ/NEMO . The activation of RIPK1 is inhibited by direct phosphorylation by TAK1, IKKα/β, MK2, and TBK1. cIAP1 was also found to mediate K48 ubiquitination of RIPK1, inhibiting its catalytic activity and promoting degradation.”

So why should you plow through all this?  Because inhibiting RIPK1 reduces oxygen/glucose deprivation induced cell death in neurons, and reduced infarct size in experimental middle cerebral artery occlusion.

RIPK1 is elevated in MS brain, and inhibition of it helps an animal model (EAE).  Mutations in optineurin, and TBK1 leading to familial ALS promote the onset of RIPK1 necroptosis

Inflammation is seen in a variety of neurologic diseases (Alzheimer’s, MS) and RIPK1 is elevated in them.

Inhibitors of RIPK1 are available and do get into the brain.  As of now two RIPK1 inhibitors have made it through phase I human safety trials.

So it’s time to try RIPK1 inhibitors in these diseases.  It is an entirely new approach to them.  Even if it works only in one disease it would be worth it.

Now a dose of cynicism.  Diseased cells have to die one way or another.  RIPK1 may help this along, but it tells us nothing about what caused RIPK1 to become activated.  It may be a biomarker of a diseased cell.  The animal models are suggestive (as they always are) but few of them have panned out when applied to man.

 

The other uses of amyloid (not all bad)

Neurologists and drug chemists pretty much view amyloid as a bad thing.  It is the major component of the senile plaque of Alzheimer’s disease, and when deposited in nerve causes amyloidotic polyneuropathy.  A recent paper and editorial casts amyloid in a different light [ Cell vol. 173 pp. 1068 – 1070, 1244 – 2253 ’18 ].  However if amyloid is so bad why do cytomegalovirus, herpes simplex viruses and E. Coli make proteins to prevent a type of amyloid from forming.

Cell death isn’t what it used to be.  Back in the day, they just died when things didn’t go well.  Now we know there are a variety of ways that cells die, and all of them have rather specific mechanisms.  Apoptosis (aka programmed cell death) is a mechanism of cell death used widely during embryonic development.  It allows the cell to die very quietly without causing inflammation.  Necroptosis is entirely different, it is another type of programmed cell death, designed to cause inflammation — bringing the immune system in to attack invading pathogens.

Two proteins (Receptor Interacting Protein Kinase 1 — RIPK1, and RIPK3) bind to each other forming amyloid, that looks for all the world like typical amyloid –it binds Congo Red, shows crossBeta diffraction and has a filamentous appearance.  Fascinating chemistry aside, the amyloid formed is crucial for necroptosis to occur, which is why various bugs try to prevent it.

The paper above describes the structure of the amyloid formed — unusual in itself, because until now amyloid was thought to involve the aggregation of a single protein.

The proteins are large: RIPK1 contains 671 amino acids, and RIPK3 contains 518.  They  both contain RHIMs (Receptor interacting protein Homotypic Interaction Motifs) which are fairly large themselves (amino acids 496 – 583 of RIPK1 and 388 – 518 of RIPK3).  Yet the amyloid the two proteins form use a very small stretches (amino acids 532 – 543 from RIPK1 and 451 – 462 from RIPK3).  How the rest of these large proteins pack around the beta strands of the 11 amino acid stretches isn’t discussed in the paper.  Even within these stretches, it is two consensus tetrapeptides (IQIG from RIPK1, and VQVG from RIPK3) that do most of the binding.

Even if you assume that I (Isoleucine) Q (glutamine) G (glycine) V (valine) occur at a frequency of 5%, in our proteome of 20,000 proteins assuming a length of amino acids IQIG and VQVG should occur 10 times each.  This may explain why 300/20,000 of our proteins contain a 100 amino acid  segment called BRICHOS which acts as a chaperone preventing amyloid formation. For details see — https://luysii.wordpress.com/2018/04/01/a-research-idea-yours-for-the-taking/.

Just another reason to take up the research idea in the link and find out just what other things amyloid is doing within our cells in the course of their normal functioning.

 

A research idea yours for the taking

Why would the gene for a protein contain a part which could form amyloid (the major component of the senile plaque of Alzheimer’s disease) and another part to prevent its formation. Therein lies a research idea, requiring no grant money, and free for you to pursue since I’ll be 80 this month and have no academic affiliation.

Bri2 (aka Integral TransMembrane protein 2B — ITM2B) is such a protein.  It is described in [ Proc. Natl. Acad. Sci. vol. 115 pp. E2752 – E2761 ’18 ] http://www.pnas.org/content/pnas/115/12/E2752.full.pdf.

As a former neurologist I was interested in the paper because two different mutations in the stop codon for Bri2 cause 2 familial forms of Alzheimer’s disease  Familial British Dementia (FBD) and Familial Danish Dementia (FDD).   So the mutated protein is longer at the carboxy terminal end.  And it is the extra amino acids which form the amyloid.

Lots of our proteins form amyloid when mutated, mutations in transthyretin cause familial amyloidotic polyneuropathy.  Amylin (Islet Amyloid Polypeptide — IAPP) is one of the most proficient amyloid formers.  Yet amylin is a protein found in the beta cell of the pancreas which releases insulin (actually in the same secretory granule containing insulin).

This is where Bri2 is thought to come in. It is also found in the pancreas.   Bri2 contains a 100 amino acid motif called BRICHOS  in its 266 amino acids which acts as a chaperone to prevent IAPP from forming amyloid (as it does in the pancreas of 90% of type II diabetics).

Even more interesting is the fact that the BRICHOS domain is found in 300 human genes, grouped into 12 distinct protein families.

Do these proteins also have segments which can form amyloid?  Are they like the amyloid in Bri2, in segments of the gene which can only be expressed if a stop codon is read through.  Nothing in the cell is perfect and how often readthrough occurs at stop codons isn’t known completely, but work is being done — Nucleic Acids Res. 2014 Aug 18; 42(14): 8928–8938.

I find it remarkable that the cause and the cure of a disease is found in the same protein.

Here’s the research proposal for you.  Look at the other 300 human genes containing the BRICHOS motif (itself just a beta sheet with alpha helices on either side) and see how many have sequences which can form amyloid.  There should be programs which predict the likelihood of an amino acid sequence forming amyloid.

It’s very hard to avoid teleology when thinking about cellular biochemistry and physiology.  It’s back to Aristotle where everything has a purpose and a design.  Clearly BRICHOS is being used for something or evolution/nature/natural selection/the creator would have long ago gotten rid of it.  Things that aren’t used tend to disappear in evolutionary time — witness the blind fish living in caves in Mexico that have essentially lost their eyes. The BRICHOS domain clearly hasn’t disappeared being present in over 1% of our proteins.

Suppose that many of the BRICHOS containing proteins have potential amyloid segments.  That would imply (to me at least) that the amyloid isn’t just junk that causes disease, but something with a cellular function. Finding out just what the function is would occupy several research groups for a long time.   This is also where you come in.  It may not pan out, but pathbreaking research is always a gamble when it isn’t stamp collecting.

 

Amyloid again, again . . .

Big pharma has spent (and lost) several fortunes trying to attack the amyloid deposits of Alzheimer’s.  But like my late med school classmate’s book — “Why God Won’t Go Away” ==https://www.amazon.com/Why-God-Wont-Go-Away/dp/034544034X, amyloid won’t go away either.   It’s a bit oblique but some 300 of our proteins contain a 100 amino acid stretch called BRICHOS.  Why? Because it acts as a chaperone protein preventing proteins with a tendency to form amyloid from aggregating into fibrils.   The amino acids form a beta sheet surrounded front and back by a single alpha helix.

[ Proc. Natl. Acad. Sci. vol. 115 pp. E2752 – E2761 ’18 ] Discusses Bri2 (aka Integral Transmembrane protein 2B (ITM2B), a 266 amino acid type II transmembrane protein. Bri2 contains a carboxy terminal domain Bri23 released by proteolytic processing between amino acids #243 #244 by furinlike proteases. Different missense mutations at the stop codon of Bri2 cause extended carboxy terminal peptides called  Abri or Adan to be released by the proteases. Abri produces Familial British Dementia (FBD) and Adan produces Familial Danish Dementia (FDD). Both are associated with amyloid deposition in blood vessels, and amyloid plaques throughout the brain along with neurofibrillary tangles.

What is fascinating (to me) is that the cause and cure are both present in the same molecule Bri2 also contains a BRICHOS domain.  This implies (to me) that possibly the segment possibly forming amyloid is being used by the cell in some other fashion.

Bri2 is found in the beta cell of the pancreas (produces insulin).  The beta cell also produces Islet Amyloid PolyPeptide (IAPP  aka amylin ) one of the most potent amyloid forming proteins known.  Nonetheless the pancreas makes tons of it, and like insulin, is secreted by the beta cell in response to elevated blood glucose.  The present work shows that Bri2 is what keeps IAPP from forming amyloid.  The BRICHOS segment (amino acids #130 – #231) is released from Bri2 by ADAM10 (you don’t want to know what the acronym stands for).

How many of the 300 or so human proteins containing the BRICHOS domain also have amyloid forming segments.  If they do, this implies that the amyloid forming segments are doing something physiologically useful.