Sunday, March 6, 2016

View Camera and a Darkroom in a Small Apartment

Although I haven't written about it on this blog yet, I have been an avid (amateur) photographer for many years. I certainly enjoy the conventional artistic side of it, composing and mindfully lighting a beautiful image. But like most things, a lot of my enjoyment also comes out of trying to deeply understand the technical aspects, getting my hands dirty doing as much of the process myself as I can, and mechanically modifying things.

I of course own a digital camera, but my favorite camera by far is this 5x4" sheet film view camera. Looks old-timey, but I actually bought it manufactured new. Some quick snapshots:



The basic principle is that you point the camera at a subject, then use knobs to move the entire lens (just that little thing on the very front) to focus, the accordion part expanding or contracting as needed to keep the inside of the camera dark. The frosted / ground glass on the back (with grid, shown) then shows a faint image of whatever you are aiming at, and you can use this to ensure focus is correct. Putting a fabric hood over your head like you often see people doing with camera like this helps see this focusing image better. Disregard the tilting and pivoting of the lens for now, I'll discuss that some other time.

Once you're happy with your focus and composition, the whole panel with the glass tilts outward as a spring-loaded clamp, and you carefully insert a film holding device (shown below) into the clamp, so that the film holder is now held in between the glass and the body of the camera.

The film holder is designed so that once the glass panel is clamping it down, the film inside the holder is now positioned precisely as far from the lens as the glass USED to be, so that if the image was in focus when looking at the ground glass, it will be now on the film too.

The film holder consists of a frame, a central membrane, then then two (usually metal) "slides" with handles at the top. The photos below show modern ones (top, closed, slides in place) and an older one but easier to see the concept (bottom). In a darkroom, you pull a slide out, and feed a piece of film (5" x 4" sheet film, roughly the size of my hand) into guides so it rests against the center membrane. It holds another on the opposite side. Then you put the slide back in, forming a light-tight seal, so it doesn't expose the film once you turn the lights back on and go traveling to a location to shoot.

Later, when you're ready to shoot, have attained focus, closed the lens diaphragm, and inserted the cartridge, you can then remove the slide, and the film will be exposed to the dark interior of the camera. Now you snap the photo. The lens itself has gears and mechanics and setting knobs on it to select an aperture and a clockwork timed shutter to expose the film for just the right amount of time you've calculated (using a light meter or just experience with lighting and weather). Then put the slide back in protecting your film. Then remove the cartridge and store it to be developed later.

(5x4" film holders, 2 sheets each Jtknowles, wikimedia, public domain)


Uwe Aranas / CC-BY-SA-3.0

This film has to be developed manually. The developing process conceptually involves only a few simple steps: go somewhere absolutely dark, remove the film from the holder, submerge it in developing chemicals for the right amount of time, rinse with mild acid to stop the developing, then submerge in a fixer chemical, which stabilizes the image and makes it light-insensitive. Rinse again, and dry, and you have a negative.

There are many interesting details to all this that I will explore in other blog posts: the science and process of how to determine your developing times and chemicals, how to test and calibrate this for a given film, where I got giant sheets of film to begin with, what the tilting and pivoting parts of the camera (80%-90% of its hardware) is for, how to digitize large negatives, etc. etc. For now I'm just going to finish up with a description of my darkroom setup, though.

I have nothing remotely approaching the space for a dedicated darkroom. I live in a small one bedroom apartment. But you don't really need a dedicated darkroom unless this is your full time job. All you need is a place you can make very dark with chemical resistant surfaces in it. For me, this was my interior bathroom, which has no windows and opens into a hallway:



Not shown is the doorway, which I sealed with double flaps of painter's tape in opposite directions--one on the frame, one on the door--to force any light to pass through an "S" curve of tape, which stops nearly all of it. I also threw a towel on the bottom of the door for good measure. It is PITCH black in there with the lights off just from that, even after 20-30 minutes of letting your eyes attempt to adjust. Darkness is actually not difficult to achieve.

Other parts of my setup:

A) You can see some film holders with film inside ready to develop on the side of the tub. It looks like I only had 4 sheets of film (two cartridges) this session.
B) Containers of concentrated developer chemical and a jug of DISTILLED water to the left. (I was not using that blender pitcher for anything other than clean water...)
C) A graduated cylinder for measuring out precise dilutions, and rubber gloves -- developer isn't incredibly dangerous, but safety never hurts. I also wear goggles, and have a fairly strong bathroom vent fan on at all times.
D) I bought a number of these convenient half liter or so plastic containers. Before turning the lights off, I pre-fill them in order with everything I need. Developer, mild acid wash (just dilute vinegar), fixer, and water, and arrange them on my tray in memorized locations. I also cut small scores into the bottles with a knife that I can feel in the dark with my thumb to identify which chemical was inside.
E) The actual film I developed in PVC tubing. The inner circumference of this tube size is perfect to hold a curled 5x4" film sheet without overlapping itself. The developed side faces inward. You can see the tube is coated in black duct tape, because even light that gets through solid PVC can be enough to mess with the film if the lights go on or the door opens (like if I need to step out to answer the phone, etc.). The pipe is capped on both ends, caps shown next to it. This long pipe holds two pieces of film in a row.
F) I have a red lightbulb on the shelf here. It turned out not to be safe for the film to have on constantly, but for parts requiring fine accuracy of pouring, I sometimes flip it on briefly. The book shields the film area from any direct light, so it's only light that bounces that arrives there. Mostly, though, I work in complete darkness with just practiced movements.
G) A simple piece of twine and some binder clips works great for drying  developed film. This is where the DISTILLED water becomes important: tap water won't screw up the chemistry, but it will leave spots of minerals on your film when it's dry, which will show up in the final image. Distilled water leaves nothing behind. The binder clip doesn't mar the image, because the very outside of the film isn't exposed--it's where the holder holds it. You can see this outline edge portion in the example photos further below.


Darkroom with low intensity red light on

Again, I'll be putting up several more posts including discussion of how to actually calculate and test developing times and chemicals and all sorts of processing steps, with many examples of the kind of results that come out, but for now, here's one example of a 5x4 developed film photo of a local hydro/gas power plant (inverted in photoshop to a positive image):

Notice I added a box in photoshop in the middle. See the top bit of that brick building in the box, with the three prongs sticking out of it with shadows cast from them? You can barely even tell it IS a brick building in the image above. But here's a higher resolution (still not full res! But you can start to see the fuzziness now) detail of a tiny bit of that brick building and one of those prongs:


Now suddenly you can review the consistency of the masons' mortaring job on individual bricks... Yeah, large format film is amazing. And this is with a bargain basement $50 or something throwaway lens, too, and film designed to be used in x ray machines, not for fine photography. It could get crisper still with professional materials and equipment. This is also a detail in the center of the image, where most lenses are good. My terrible one I used here gets very soft at the edges, while a high end lens would not.

One more example photo:



Tuesday, December 8, 2015

Backstrap Loom Weaving

Hello again everyone! I am recently finished with my doctoral training and in a new job and back to crafting again. Well... technically, this post's topic of weaving I originally started as a stress-relieving activity during the height of my studies. Weaving is meditative and creates satisfying, well-ordered structure in front of you with minimal mental gymnastics. At least if you're not making any complex fabric.

"Weaving" probably initially brings to mind images of gigantic, unwieldy, wooden looms taking up half of one's living room. And indeed, if you want the height of industrial efficiency, stylistic freedom, and ease of creating different finishes and textures, that's the way to go. But I was A) poor, B) didn't have 100 square feet to spare, and C) just wanted a relaxing craft activity. So I chose to use backstrap weaving techniques:

Backstrap weaving (image by Infrogmation, GNFL license)

Backstrap looms, like the one shown above, are cheap to make, requiring only cords, straps, a series of wooden dowels and slats, and some spare lengths of the same kind of yarn used to weave. The fabric you are making itself holds the loom together otherwise. When not in use, the whole project can be rolled up and stored. While in use, the loom is held together by placing tension on the warp yarns (vertical to the weaver) with your body, leaning back and stretching them between yourself and an anchor point in front of you. Dowels then keep the yarns separated while you weave the weft yarns (horizontal to the weaver) back and forth between. As you finish sections, you partially roll them up on the bar nearest you so that the next section is in reach, and so on until finished.

I didn't start with anything as ambitious as the woman above's yard-width cloth. I started by making a simple, thin, acrylic yarn strap. Below is a third party technical diagram of a backstrap loom, followed by a photo of my actual homemade one:



The warp threads are actually one continuous loop back and forth between the two anchor points of the loom, at the body and the far end. The warp threads can be seen splayed out at (A, furthest from the weaver) in the image. These tend to get tangled on their own, though, so to help prevent this, I made a "cross" (B) in the yarns by weaving a simple stick between every other thread, then weaving another one between the opposite threads (actual cross is between the sticks at B). This forces any tangles to resolve themselves and leave the yarns in order and well-spaced going toward the working section of the loom. To help keep the yarns spread out and untangled, tension is applied by the bar near the body (C), which is attached to a strap around the weaver's back (not shown, though notches are visible for the strap to attach). Two sticks are used here to allow the finished work to be rolled up as progress is made, without unrolling from the tension.

Then the actual process of weaving involves taking another length of yarn, the weft, and passing it through all the odd numbered yarns, then all the evens going the other way, back and forth. The weft thread is all rolled onto a bobbin (D, I used a knitting needle) so that passing all of it through the warps doesn't take an hour each time. The working edge of this piece of cloth is at (E); you can see where the weft has made it to so far. The stick right at the working edge is used to "beat down" the weft toward the finished cloth after each pass to make it nice and tightly woven.

How does the weft get between opposite yarns each pass? Well, you could thread it through each time manually, but that would be horrible. Instead, there's a system to quickly shift all the even numbered yarns up and then all the odd ones, quickly and efficiently. The odd numbered yarns are held apart by a nice thick bar (F) at the back of the loom. The even numbered yarns are then each individually tied to little strings ("heddles") that reaches down through the odd yarns, and then attaches above to a stick (G). To raise the odd yarns, the thick bar (F) is moved back and forth and up and down to well separate those yarns, while the heddle strings allow the even yarns to pass below without getting in the way. This creates a triangular space between (the "tent"). The beater stick (E) is placed inside (it is removed and replaced each pass), and flipped on its side to make the tent even bigger. Then the weft bobbin (D) is passed through. The weft is beaten down, and the beater stick removed again. Next, the thick bar (F) is pushed back out of the way, and the stick with all the heddle strings on it (G) is lifted, PULLING all the even-numbered strings through the odd ones so that they are now on top. The beater stick is put in again, the weft bobbin passed the other way. These two phases are repeated over and over again until a cloth is formed.


Above is the finished result of my first strap. If you look very closely, you can see several characteristic errors made by people who don't know what they're doing. For one thing, one end is about half as wide as the other. This is called "pulling in" and it results from not leaving enough slack in the weft thread as it is woven through the warp. When it gets beaten tight, the weft is crimped, and thus shortens. If there's no slack, this pulls in the sides of the whole cloth, making it progressively narrower. The way to fix this is to leave the weft at a slight angle before beating it tight, so that it is a bit longer than the width of the cloth, leaving room for crimping.

You can also see that at the edges, the texture changes in places. It becomes more square looking, while the middle of the strap looks hexagonal. This is because the warp yarns at the edges were looser than the ones in the middle. I didn't have even tension across my whole width. Thus, the warp and weft at the edges are about equally tight and pass over one another equally, while the warps in the middle are much tighter than the wefts, so one of them crimps and the other doesn't, leaving the hexagonal look as you can only see one of them. On the sides you see both warp and weft.  My next attempt was a lot better:



Here I was making a small patch of a twill fabric (specifically a gabardine). The first strap I made was a "plain weave", over under over under. In a basic twill, instead of having all the odd yarns alternate with the evens, what you do is pass over TWO warp yarns, then under two, then over two... Which two you pass also shifts over every time you send the weft through, like this:

Twill Weave (by Jauncourt, CC Attribution Share-Alike)

If you imagine counting off every warp thread as 1,2,3,4,1,2,3,4... this means that depending on the pass of the weft, you need to variably lift all the 1's and 2's, then next all the 2's and 3's, then 3's and 4's, and finally the 1's and 4's. Four different sets of yarns need to be controlled instead of the two sets for the simple strap I made. This means that I need not just one thick bar and one set of heddle strings, but instead one thick bar and THREE sets of strings. This way, I can achieve all four types of lifted sets of strings. You can see the three different tied sticks in the photo above. This is trickier to set up, but not much trickier to weave. I'm using orange yarn for the warp this time, and red for the weft, so that you can tell them apart later.


Here's one side of the finished product (above). Notice that you can ONLY see orange. This is a "warp-faced" fabric since on the intended display side (above), you can only see the orange warp yarns. The red wefts are hidden underneath. Notice also that the cloth looks much denser and tighter than the strap earlier. Since the yarns aren't crimping as much as in a plain weave (only every two yarns not every one), more yarns are packed into a square inch than in a plain weave. This makes twills harder-wearing and popular for work clothes and jeans. The tight weave also makes twills more insulating (jeans are generally warmer than linen pants, which use plain weave), and easier to waterproof. Gabardine, the weave I made, is popular as an outer weave for raincoats, for instance.

Here's the back side of the same cloth. On the back, you can see the warp and weft about equally well. This is not the "face" of the cloth, though, so it's still called "warp-faced." Notice that the width doesn't change quite as dramatically as last time, but I still pulled in a bit from start (left) to finish (right). The tension is more even this time. The edges are also more consistent, from repetitive practice.

Next, I used a mixture of stiff hemp warp strings and soft acrylic weft yarns, and chose a new pattern: a diamond twill. This is a normal twill, but with the pattern simply changing direction in blocks. When tying the warp strings, I changed the pattern partway, and I also changed the direction of which heddles I lifted occasionally while weaving to make the pattern change the other way as well. Here's the loom set up first, followed by a diagram of a diamond twill.


If you look closely, this looks just like a normal twill but in smaller "blocks" that flip occasionally.

Again, this pattern required 4 types of yarn lifts, so a bar + 3 heddles, varying which sets of warp yarns were lifted as I went. The finished product is below, with a bit fancier finishing of the ends. You can see the diamond pattern, though it's not a consistent looking as I'd like. You can see both hemp and red yarn here, because even though it's warp-faced, the yarn is so much bulkier that this balances out the tension visually and keeps them both more equally visible at once. Notice that when I cut the bottom, it by no means immediately unravels. These cloths could definitely be cut to patterns and used in sewing if large enough.


It takes me about 20 minutes to wind the warp yarns around two pipes in a wooden board and transfer them onto the loom, then anywhere from 30 minutes to hours to tie the heddles depending on complexity and width of the cloth, and about 5 minutes per inch of cloth to weave the weft along. Next up is going to be attempting a much larger piece of fabric, like a dish towel, perhaps. I also want to try out changing the warp and weft colors across the fabric, such as when making a plaid fabric.

Much fancier and irregular patterns can also be woven into a cloth, like the shapes of animals, but this requires manually "dropping" or "picking up" yarns in violation of the background pattern as you go, and it is very painstaking by comparison to plain fabrics like these.

Wednesday, March 18, 2015

Where is Mr. Brassica?!

Hello Cauliflower Lab readers!

I am Mr. Brassica's assistant. I wanted to let everyone know that Mr. Brassica has had to put his blog projects on hold for the spring in order to complete IRL experiments and publications.

He will be back in May or June with new posts.

Thank you for your understanding!

-- A.P.

Tuesday, December 2, 2014

Making Potash



"Potash" is a vague term referring to any of a variety of soluble compounds of alkali metals, mostly potassium carbonate, potassium chloride, sodium carbonate, or sodium chloride. All of these compounds can be found in burned plant ashes as a portion of the leftover non-flammable, non-volatile material left after burning.

I desire potash for one reason: it acts as a flux for silica compounds. I will use this in two different other projects. First, as a glaze for pottery that can melt and seal ceramic items without the underlying item melting, thus allowing me to waterproof earthenware mugs and pots. Second, to lower the melting point of quartz for making glass at reasonable temperatures.

When you burn things to create ashes, though, a lot of other substances also end up in ashes:
Iron oxides
silicon oxides
aluminum oxide
calcium oxide
magnesium oxide
manganese oxide
phosphorus oxides
unburned charcoal

The ratios of these and potash depend on the species burned. Willow wood yields a whopping 50% potash, and very little silicon or aluminum. English oak yields 23% sodium and potassium compounds. Rice husks only provide 1.3% potash, etc. I am using leftover ashes left over after other clay project firings, and I don't know the species. This is bad for achieving a consistent pottery glaze. Later, I will want to source a consistent supply of ash, but right now I'm happy with ANY result.

The best way to separate our desired potash from the other substances in raw ashes is to exploit potash's solubility. So first, we shovel some ash into a container, (I strained through a cheap spaghetti colander first to get rid of large chunks), flood it with water, and let it sit for a couple of days. It will look like this:


In the photo, the liquid on top is full of potassium, sodium, hydroxide, and carbonate ions, which we can now pour off into another container for further processing. The solids at the bottom are the non-soluble oxides that would ruin the melting point-lowering properties of the potash.

Don't get greedy. Don't pour off the liquid before the solids have settled, and stop pouring when you start getting any ashes pouring out. The non-soluble ashes will be counter-productive, because they are highly refractory and will raise the melting temperature of whatever you're making.

Careful, the poured liquid is highly caustic. It is not a cellular poison in the sense that something like cyanide is, but it will chemically "burn" you (turn your skin into soap) if you get it on you and don't wash it off quickly. Wear eye protection and gloves when pouring, and flush any splashed skin with large amounts of water. Treat this liquid as essentially Drano.

By the way, save the byproduct solid ashes. For two reasons: One, you can refill with water and repeat the process again for more liquid from these several times. Two, once you're all done, the remaining washed ashes are wonderful for making fire bricks for furnace linings, because all the remaining materials are very resistant to heat and have very high melting temperatures.

Anyway, back to the liquid: letting this air dry is ineffective, because crystals will start covering the water surface and prevent evaporation. Boiling the water off works better. The amount of water you see at the top of the jar poured into a pan and boiled dry yielded this:


Again, these crystals are caustic. They will re-dissolve and saponify your skin if you get them on you with any moisture (like sweat). Be careful. That said, I unceremoniously scraped mine off with a spatula. The potash won't ruin your utensils or pot, and you can use them for food later, as long as you VERY thoroughly rinse and wash the potash off first. Tiny trace remnants that remain after going through a dishwasher or a good scrubbing will not hurt you, they are not toxic beyond being strong bases. Think of it like having Drano or bleach in a pot temporarily -- you wouldn't throw out the pot, but you would wash it well before cooking with it.

At the end of the day I disappointingly ended up with this much potash from one pouring of the container liquid:


These aren't exactly fluffy white crystals, but that's okay. It might be just that there are several compounds mixed here, so they inhibit one another's crystal formation. Since all of the hydroxides and carbonates are useful, though, this is not a problem. We don't need to separate or purify them individually (also, the crystals are still damp in this photo).

Don't put the potash on aluminum foil. It will react.

This is only one pour off the jar of ashes. You should be able to continue getting potash for many pours in a row, leave a day or two between each one, until you stop getting justifiable returns. My second pour-off a couple of days later yielded about 75% as much as the first, for example.

The potash will lower the melting point of clay it is mixed with. So when mixed with crumbled dry clay, ground up, mixed with water, painted onto a piece of ceramic, and left to dry, the surface mixture will melt sooner than the underlying clay piece, creating a glaze. I will cover this topic in more detail in a future blog post, but I have already achieved early success in using my potash to make glaze. The black corner of the right side piece in this photo is vitreous glaze made with the above crystals and clay powder. If/when I can cover a whole piece with successful glazing like this, I will be able to make waterproof items:



Sunday, November 9, 2014

Homebrew Pottery - Firing local clay



SO! I have now found some local clay and I have figured out the best amount of tempering to use. I also mixed up my pure clay with some different additives for the purposes that I wanted.

Now it's time to fire the clay into true ceramic! Several processes are involved, but mainly, the firing process mainly drives out water (only some water evaporates, and the rest is physically trapped or chemically bound), burns off organic material, changes the crystal structure of quartz components ("quartz inversion"), and sinters particles together.

My "kiln" at the moment is nothing but a Weber grill. The temperatures on the rack where you would normally put food only get up to maybe 600-800 degrees, but by completely burying clay wares in charcoal, the temperatures can get up to the roughly 1800+ degrees needed to begin to make true, low-fire earthenware ceramics (bisque temperature). The whole heating process needs to proceed SLOWLY. For more information, google "firing schedules." Details are beyond the scope of this blog, unless people are interested.

I'm firing simple bricks for now, to be used later as kiln components for a proper homemade ceramic kiln. The orange bricks in the photos are ones I fired earlier, and have painted with a homemade ash glaze as test pieces for glazing. This is simply part potash and part local clay, mixed with water and painted onto the pieces then left to dry in place. You can see the grayish ash on top of them (see future blog posts for more details).

First, I heat up the bricks around the outside of the grill, as the coals are lighting. This very gently brings them up to ~150 degrees over the course of half an hour to an hour, driving out as much water as possible without creating expanding steam that might crack the pieces:


Next, I move the bricks into hotter and hotter zones, eventually leading to a pile in the middle. This brings them closer to maybe 400 degrees, driving out the last bits of chemically bound water.


I remove the grill and very quickly (lest the pieces start cooling off) add more fresh charcoal to the fire to make a nice bed. I pile the pieces back on top of the new charcoal bed and leave them there for maybe another half an hour or so. You can see in one of the photos below that open flames are visible from organics burning off, and the bits of grass are charring away. this is somewhere between probably 600-1000 degrees. 

I did this part too quickly still and caused some black charred-looking areas, which is a sign of too little oxygen and too quick of organic fuel consumption (it has to do with the iron in the clay body reducing), but oh well, lesson learned for next time.




Once most of the organics looked burned off, I covered the pieces entirely with a layer of lump charcoal. This decreases oxygen access, so AS SOON AS POSSIBLE they need to have a forced airflow added to them at this point, or further "black coring" will occur, which is bad for your clay. I had a box fan pointed at the fire when I did it at this point, but this was probably not enough.


The upper charcoal is lighting, and you can see the red-hot core of the inside part of the grill where the pieces are. Open flames are still visible, but at this point I wouldn't be too concerned about it being organic material, it's probably just the charcoal itself. This is dull to cherry red glow around 1500 degrees maybe at this point:


The firing continued into the evening hours. The internal temperature is now sufficient to create bright red to orange hot ceramic pieces inside, as you can see in the photo where I briefly lifted the lid off to show a brick inside. This bright orange to almost yellow glow is probably somewhere around 1800 (+?) degrees. During this time, I was blowing air into the fire from a squirrel cage electric powered fan, positioned underneath through the ash collection hole, firing oxygen up through the fire. I got lazy and was inconsistent with this, but it really should have been going continuously, all the way until the fire burned down, if I wanted the strongest pieces. Since I let it go without forced air for awhile, it further reduced more of my clay's iron and made the pieces more brittle than they should have been.



This temperature was held for maybe another hour or so, adding charcoal as necessary. After that, I let it die down on its own all the way until cool (several hours). Then I removed the bricks and took the photo you see below. The bricks have been successfully made into true ceramic, with their oxidized iron components giving them a familiar brick hue. The especially light color is no doubt due to the sand mixture I added.

The gray/black spots you see are reduced iron from me being impatient with my firing schedule and not using enough forced oxygen. These reduce the strength and effectiveness of the bricks and are to be avoided.

Generally, the bricks were very brittle, probably too brittle to use. Not just from black coring, but from probably using too much sand and grass clippings in the mixture. Just from touching them, it is OBVIOUS though that they have huge insulation value. The bricks are very lightweight and very porous from the burned away grass, almost like pumice. All those tiny air pockets make them more effective at insulation, at the expense of fragility. These need a higher percentage of clay overall, because the balance is too far toward fragility, but it's a minor recipe tweak.


The other three pieces were ones that were fired earlier, and simply had glaze applied to them. You can see how they hold specific shapes much better and have little or no porosity, because these pieces were 100% clay in composition (on the inside/main body). You can also see how they are redder in color due to higher iron content when sand and grass are not mixed in.

The glazing experiment was a partial success. From left to right in the image below, glaze mixtures with a lower ratio of clay slip to potash are shown. The highest clay content on the left strangely seemed to have burned away completely, or perhaps got scraped off by accident, this one is confusing. The other two are as expected: the medium mixture in the middle looks like I just painted more earthenware right onto the block. This is no true glaze, and won't help to keep water out, etc.  But the high-potash content piece on the right shows signs of JUUUUUST beginning to form an actual glaze in the upper left, and a near-glaze (white) elsewhere. The black corner is actual glass, and is proper, waterproof glazing!

Getting the whole piece to be consistently covered will require a combination of just a slightly thicker coat of glaze painted on, a couple hundred more degrees of temperature, and probably more potash in the recipe. The more potash you add, the lower the glassifying temperature, but the more fragile and likely to crack (and harbor bacteria) the glaze is, so this is a careful balancing act, but I feel confident now that an actual waterproof glazed mug is within my abilities!


Next steps include making a more proper kiln with insulative lining to get hotter temperatures, adding better airflow and oxygen control throughout the process, and fixing the glaze and brick recipes.


Monday, October 27, 2014

Geology Simulator -- Minerology and Petrology Systems


In this post, I outline how the world will keep track of rocks, minerals and their properties, and why. I had to take a detour from plate tectonics to at least think about this issue, in order to make certain early database decisions.  And by "think" I mean "read several textbooks" (well no not yet, but I'm most of the way through one!):



Unlike a game like Minecraft, I don't want to store my material information in terms of only distinct categories with a small set of ID numbers. Whenever possible, I want to store CONTINUOUS information about things like chemical makeup, so that rocks for the most part vary smoothly and gradually across the world. ...Unless there's some reason for them not to, like in the case of a sudden magma tube or a fault line.

If I DID keep track of a fixed number of categorical rock types and then looked up their chemistry, it would lead to sudden, jarring jumps in rock types out of nowhere. These sudden jumps would then get worse over time as plates moved around and erosion occurred. It would also cause chemicals to pop in and out of existence, because if you change from potassium rich categories to potassium poor, the potassium just goes poof, which violates conservation of matter. If instead I keep track of element and mineral concentrations and then look up what type of rock it is if and when I need to name it, these problems don't happen.


These are all granite! 
Yet as you might imagine, they have very different properties,
ores, etc. One named category covers a lot of continuous ground
that I don't want to miss out on!
There are a couple major problems with this continuous approach, though. One, it adds up to a lot more memory space than a cheap little ID number (Minecraft style). I've run some test simulations that simply populate the world with goxels that have all the information, and it was already pushing 1.5 gigabytes or so with just a medium size world, just sitting there (I'm trying to stay within 2 gigs for 32 bit users).

The other problem is even more pressing: I don't know how to guess a material's hardness, for example, from scratch (unlike chemistry, see later in this post). Hardness is an example of a number that depends on complex things like crystal structure and bond strengths that I couldn't possibly keep track of reasonably.

To kill both birds with one stone, I will put things like hardness in pre-supplied text files for index rocks (that I look up in books), then on program startup, interpolate between those reference points for rocks in between and store this in lookup tables. Later I can go grab the interpolated numbers during simulation whenever I need them. So we have two types of data: the goxel and the lookup table type:

DATA DIVISION 



Data Type 1, Information we can carry around in each goxel
  • Whether it is a rock at all -- goxels can be rocks, or other kinds of things like liquids, gases, biological data holders, sediments, and so on. Those things reinterpret the data below as they need, instead of wastefully keeping additional space in memory that would usually be empty.
  • Rock class -- igneous extrusive and intrusive, sedimentary, metamorphic.
  • Rock subclass -- clastic, chemical, and biological sedimentary rocks, and metamorphic rocks that keep track of their different original rock sources.
  • Chemical makeup -- relative proportions of ~20 geologically relevant elements in the goxel.
  • Texture -- i.e. whether the rock is made up of finer or coarser grains (several levels)
  • Highest temperature reached in the past (for metamorphic rocks)
  • Highest pressure reached in the past (for metamorphic rocks)
  • Current temperature
  • Current pressure
  • How much rock there is left, out of one full goxel -- keeps track of erosion having removed material.
  • How cracked and broken down the rock is -- this accelerates weathering, and acts as a path for aquifers and mineral veins and magma, etc. Also, if it reaches the maximum value (gravel sized bits), the goxel becomes a sediment goxel.
  • The identities and proportions of the few most prevalent rock-building minerals (special subset of minerals) that we have determined are currently in a solid goxel.


by Marie-Lan Nguyen

Data Type 2, Information we look up from centralized reference sources only as needed
  • Name of the NEAREST rock category for human flavor (not geologically important)
  • Melting point
  • Chemical resistance -- to weathering
  • Hardness -- grinding resistance
  • Mechanical resistance other than hardness -- bending, pulling, etc.
  • Thermal expansion coefficient -- influences cracking in temperature cycles.
  • Porosity -- High porosity creates vulnerability to frost and salt wedging even in solid rock.
  • Density -- affects how heavy a goxel is for floating on the mantle.
  • Specific heat -- affects thermodynamic calculations, of course.
  • Individual properties of constituent minerals.
Most of the things in Type 1 are used often, can't be categorized validly, and can plausibly be condensed down into compact shorthand formats. Temperature, for example: we don't need to know the temperature of everything down to a degree. So for example I can actually store temperature more like just "steps of 100 degrees C at a time, covering only ranges relevant to rocks melting."

On the other hand, the things in Type 2 are there because they are used rarely and/or they change nonlinearly or in complex ways. But what do we do if we need to know a goxel's melting point right now? We use lookup tables to figure it out on the fly!



ROCK IDENTIFICATION AND INTERPOLATION 

The basic idea here is to choose the few critical variables for classifying rock types from amongst the stuff carried around by goxels, then make big tables of all the different possible combinations of those variables that hold the other data we want, just like we want it, ready to quickly read off.

Actual, known rocks are read from user text files (there will be defaults) placed at their respective reference positions in the tables. Then the remaining spots are all interpolated between their reference neighbors. So a spot 1/3 of the way from granite to diorite will store values averaged between those on file for granite and diorite, biased toward granite.

This is all calculated once at startup, using user-customizable entries in a text file for the reference rocks. This means that if you want to add your own rocks, you just say where, and the algorithm will automatically blend those properties into the local neighborhood when you start the program up. Then your new rock type will appear exactly where you would expect it to in your world at the very end, procedurally.

We can't just have one big dumb lookup table for everything, because it would take up like 20 gigabytes or something. But that's okay, because many dimensions are mutually exclusive anyway. Rocks formed in different environments can be delegated to their own smaller tables. Basically we end up with a flowchart of smaller tables that add up to only small peanuts of storage:

Some example reference rocks are listed by number on the chart above. You do need one rock type at least for each box, but after that, it's up to you how densely to fill in the space. For example, let's consider three versions of just the Igneous lookup table, with different reference rocks:
If you only have granite in that category in the custom text files, the game will treat all igneous rocks as identical to granite and call them all granite at the end. If you have granite and basalt, it will name some rocks one and some the other, with the dotted line showing the naming cutoff. The smooth color transition represents in-between rock material properties averaged across the space for in-between goxels. Adding in andesite makes the space more complex, and so on.

By having these tables precalculated, looking up information is very quick. Find your table, then look up your exact coordinates and just read the data off. There is no need for calculating distances or averaging any values during the actual simulation. The time saved this way should make up for the time calculating the tables many many times over, and goxels don't have to store as much info, either.

...So that's how we look up whole rocks, but I also mentioned minerals. What are those, exactly, and why do they also matter?



MINEROLOGY -- VERY BASICS

Rocks are usually just mixtures of a finite set of minerals. Minerals are consistent substances with specific chemical formulae that will tend to form crystals or other reliable formations. Granite has no chemical formula. If you look at a chunk of granite, you can clearly see different distinctly colored regions. Each of these is a coarse grained crystal of one of a variety of pure (or at least homogenous) minerals:

by Friman CC-BY-SA-3.0

In the case of this piece of granite, the pinkish crystals are probably some sort of feldspar, the white ones are quartz, and the black ones are either amphibole or biotite minerals. If you look much more closely, the crystals become quite striking in how distinct they are (not the same sample):

by Thomas Bresson CC-BY-SA-3.0

Why are these minerals so separated? Because granite is a rock formed by very slow cooling of magma trapped beneath the surface. As the magma passes gradually through different temperature ranges, specific minerals have a lot of time to crystallize out of solution all by themselves, because hotter crystallizing minerals already left solution and cooler ones aren't in range yet. And its different minerals don't blend very well, because they have different crystal structures. So all the crystals of each type end up relatively large and distinct.

By contrast, a rock that cools very quickly does not have visibly sized crystals, or has very small ones. This rhyolite rock below is probably almost chemically identical to the pink granite shown above, but cooled much more quickly (at the surface, in an eruption):

by Michael C. Rygel CC-BY-SA-3.0

Do I really care whether rock looks coarse or fine? Well no, I don't care how they LOOK, because I'm not modeling visual textures. But I do care VERY much about crystallization and minerals in general, for a few reasons. All of the reasons revolve around the same theme of minerals influencing how chemical elements move around in non-homogenous ways:
  1. Different minerals are more or less stable against erosion and weathering. Generally, heavier minerals that form deeper down (like olivine) tend to be less stable at the surface than lighter minerals that form closer to the surface (like quartz). So olivine will weather away into small clays quickly, and quartz will weather more slowly into larger grains of sand. These then get separated out by wave actions, for example, and form their own entirely different types of sediments (shales and sandstones) in physically distant locations. Entire landscapes depend on different minerals eroding away at different rates!
  2. If crystals are already tinier than sand, then they won't probably erode into sand, for example. This compounds with #1 above.
  3. Crystallization of minerals from magma (and melting into magma in reverse) can have profound consequences for distribution of chemistry in the planet. Crystals can sink in their magma as they form, as just one example, which can concentrate heavy minerals lower down and concentrate lighter ones on top. This is no less important than the main driving force behind continents forming and persisting over eons!
  4. Metamorphic rocks are partially a result of the shape and texture of rocks changing with temperature and pressure, but also partially a result of constituent minerals destabilizing and re-crystallizing in new forms with depth and heat. The minerals, not just a bag of loose elements, matter.
The erosion parts (#1-2) are pretty simple. Erosion just simply pays attention to minerals and then moves sediments in a biased fashion based on their properties, no big deal. The metamorphic issue (#4) I am going to ignore for now, because metamorphism isn't as important early on in my project. The crystallization (#3 above) is a lot more complicated and also important sooner, for igneous rocks:
IGNEOUS CRYSTALLIZATION

Crystallization can happen in a lot of different ways, and depending on which way it happens, it can make a huge difference for geology. There's three main variables as far as I understand it so far: 
  1. How many substances are present
  2. Hwo well they form a solid solution. This means a solid that can substitute ions out of its structure somewhat freely. Alloys of metals are solid solutions, but mineral crystals can be too. Usually this happens when two minerals have similar crystal structures except for a substitution of one element... often in the same element series, such as calcium <--> magnesium. The different ions can diffuse in and out and essentially "dissolve" two types of crystals together. If a solid solution doesn't form, you'll get different distinct crystals next to each other instead.

    Solid solutions are on a continuum, based on how WELL ions diffuse in a given situation. From perfectly, to not at all. Also, what matters is ACTUAL diffusion. Rate of cooling or density of crystals can get in the way even if a solid solution / diffusion would have occurred eventually otherwise. So "low diffusion" can mean "chemically won't ever diffuse" or it can mean "would if it could but cooled too quickly" or it can mean "Crystals sank and aren't physically there anymore to diffuse into even if they would have qualified."
  3. Whether each crystal is heavier or lighter or the same density as the remaining liquid after crystallizing.
These can combine in different ways to yield very different outcomes. Below is a very simple (I know it doesn't look like it!) phase diagram and some outcomes from it in different situations:


Okay, so the graph is a diagram of how two substances will crystallize (or melt in the other direction) based on temperature in different mixtures. On extreme left or right, you have pure substance A or B, and melting or freezing happens at a single point: the melting points. In the middle, you have a slurry phase, where the substances are crystallizing together not all at once.

Given example initial composition M, as the temperature lowers: T(ime)0 will be pure liquid. T1 is when the first crystals begin to form, where it intersects the "liquidus line." At T2, we are in slurry form. At T3 OR T4 (depending if A and B can form solid solution together), everything has crystallized.

The circles lower in the diagram outline different chemical situations:

The top situation 
...has good ion diffusion and equal densities of everything, and the solid is a solid solution. Crystals are distinct temporarily, but eventually, ions diffuse evenly through the substance and the final solid is homogenous again, just like the liquid. (Sapphires are another example of this -- iron "dissolved" in a solid solution of corundum)

The middle situation 
...shows low solid diffusion and crystals that sink, but the materials can still form a solid solution at freezing time. As you can see, you end up with a "sunset" chemical gradient from the top to the bottom of the lava chamber. This is largely responsible for continents existing: the lighter weight crystals are less dense and concentrate on top, and thus top rocks float better on the mantle than the rock they came from.

The bottom situation 
...shows non-solid-solution (immiscible) crystal forms that do not sink or float very well, and it forms a patchwork of different solid crystals that looks like, say, granite.

The phase diagrams can get more complicated. In the following diagram, there is a solid mixture that is only a solid solution on the sides, and converts to distinct crystals in the middle., depending on concentrations of each substance and temperature. You also end up with two different sets of slurry curves, and a center "eutectic point" where both crystals form at once:


You can also have situations with more than two substances present, in which case the curves become planes instead, and you have more of a triangular prism (in the case of three substances here):

Start at some triple composition of A, B, and C near the top. Cool down gradually, following
a path as if a ball were rolling down the hills. When on one planar surface, only
crystallize that substance. When on a middling ridge, crystallize two. 
At the bottom point (Eabc), crystallize all 3 until solid.

Add in a whole bunch more materials, pressure as a variable, and the need to use shortcut formats since you can't draw an 8-dimensional space, and it gets almost incomprehensibly complicated and difficult to interpret by eye:





SIMULATING CRYSTALLIZATION

The good news is that I don't think all this is actually that hard to code, at the end of the day! 

First of all, I can simply restrict the crystallization calculations to be limited to 3 substances, so a triangular prism is the most complicated we can get. The goxels keep track of their 3 most prevalent minerals to aid in this. This is good enough for tons of realism.

Second, there are a limited number of common situations that occur. I showed you a simple binary phase diagram, and a "eutectic" diagram (the two-lobed one), for example. There are only 8 or so normal types of diagrams like these, all similar to one another. I could code explicit instructions for each, without having to have a program interpret diagrams procedurally. I'll actually probably start out with just the 2 or 3 most common situations, and even that might be good enough.

Third, the program doesn't have to consider full 2-D or 3-D diagrams. It only needs to know the equations for the curves or planes, because that's where things actually matter. Thus, all this data takes up very little space and it is actually pretty quick and easy to work with.  The rules are also the same for 1,2, or 3 dimensions, just pasting in a Y or Z variable as needed basically (sort of like a Euclidean distance equation).

Fourth, I don't need to have actual correct data for curves!! I will want to know which minerals actually form solid solutions and so on, but getting the actual curves exact is totally unnecessary. I can just encode the true melting points of different substances, and then have the program guesstimate a reasonable "almond shape" for the slurry phases, and nobody will be able to tell. I'm confident the geology will look fine at the end, at least that's how I'm going to write draft 1.

Fifth, I don't even need curves at all. I can estimate a curve as even just 2 or 3 flat lines or flat planes, and it will probably be good enough, and much faster to calculate and code up. So instead of almond zones, we will have polygons.

AND this diagram is only for your benefit.
The program just sees it as a list of simple information, such as:
0% B, slope = 0.7, liquidus
30% B, slope = 0.6, liquidus
70% B, slope = 0.1, liquidus
0% B, slope = 0.2, solidus
30% B, slope = 0.65, solidus
70% B, slope = 1.3, solidus
Melting point A = 21
Ion diffusion = 45%
Density A = 16 solid, 14 liquid
Density B = 20 solid, 15 liquid

Any elements / chemicals that get left out of the crystallization calculations? They are treated much more simply. For example, heavier ones might tend to sink a bit between goxels and lighter ones float up, but otherwise, whenever the crystallization is done, leftovers just get locked in, in an abstract sense, in that goxel. Rare ores and gems can wait to be calculated until the end of simulation.

NEXT

Anyway, that was a LOT of information! And still no demo code! This stuff is going more slowly than I hoped, and I keep running into side issues, but the tectonics and the functional demo march ever closer even in the meantime.

In my next post, I promise I'll have a down and dirty demo of simple tectonics in action!
(Update: Promise completely broken, but I do still plan to get back to this.)