Presentation is loading. Please wait.

Presentation is loading. Please wait.

Chapter 24. Stable Mineral Assemblages in Metamorphic Rocks

Similar presentations


Presentation on theme: "Chapter 24. Stable Mineral Assemblages in Metamorphic Rocks"— Presentation transcript:

1 Chapter 24. Stable Mineral Assemblages in Metamorphic Rocks
Equilibrium Mineral Assemblages At equilibrium, the mineralogy (and the composition of each mineral) is determined by T, P, and X “Mineral paragenesis” refers to such an equilibrium mineral assemblage Relict minerals or later alteration products are excluded unless specifically stated “Mineral assemblage” is used by some as a synonym for paragenesis, conventionally assuming equilibrium for the term Impossible to prove that a mineral assemblage now at the Earth’s surface represents thermodynamic (chemical) equilibrium at prior elevated metamorphic conditions Indirect textural and chemical support for such a conclusion is discussed in the text In short, it is typically easy to recognize non-equilibrium minerals (retrograde rims, reaction textures, etc.) We shall assume equilibrium mineral assemblages in the following discussion (will ignore retrograde…)

2 The Phase Rule in Metamorphic Systems
Phase rule, as applied to systems at equilibrium: F = C - f + 2 the phase rule (Eq 6.1) f = the number of phases in the system C = the number of components: the minimum number of chemical constituents required to specify every phase in the system F = the number of degrees of freedom: the number of independently variable intensive parameters of state (such as temperature, pressure, the composition of each phase, etc.) A typical sample from a metamorphic terrane Likely select a sample from within a zone, and not from right on an isograd Alternatively, pick a random point anywhere on a phase diagram Likely point will be within a divariant field and not on a univariant curve or invariant point The most common situation is divariant (F = 2), meaning that P and T are independently variable without affecting the mineral assemblage In complex natural systems there may be one or more compositional variables as well, so that F may be greater than two The common occurrence of certain metamorphic mineral assemblages worldwide supports this contention that F  2, because such assemblages are much more likely to represent variable P-T-X conditions than more restricted situations

3 The Phase Rule in Metamorphic Systems
If F  2 is the most common situation, then the phase rule may be adjusted accordingly: F = C - f + 2  2 f  C (Eq 24.1) Goldschmidt’s mineralogical phase rule, or simply the mineralogical phase rule The above simplified phase rule states that in the most common situation for a rock at equilibrium the number of phases is equal to or greater than the number of components It has been called Goldschmidt’s mineralogical phase rule, or simply the mineralogical phase rule It is useful in evaluating whether or not a rock is at equilibrium

4 The Phase Rule in Metamorphic Systems
Suppose we have determined C for a rock Consider the following three scenarios: a) f = C The standard divariant situation The rock probably represents an equilibrium mineral assemblage from within a metamorphic zone To some authors the mineralogical phase rule is phi = C

5 The Phase Rule in Metamorphic Systems
b) f < C Common with mineral systems that exhibit solid solution Plagioclase Liquid plus For example, in the plagioclase or the olivine system we can see that under metamorphic conditions (below the solidus) these 2-C systems consist of a single mineral phase phi = 1 and C = 2 so phi < C

6 The Phase Rule in Metamorphic Systems
c) f > C A more interesting situation, and at least one of three situations must be responsible: 1)  F < 2 The sample is collected from a location right on a univariant reaction curve (isograd) or invariant point The full phase rule (F = C - f + 2 ) indicates that, if C is constant, for every decrease in F, phi increases by one

7 The Phase Rule in Metamorphic Systems
Consider the following three scenarios: C = 1 f = 1 common f = 2 rare f = 3 only at the specific P-T conditions of the invariant point (~ 0.37 GPa and 500oC) Figure The P-T phase diagram for the system Al2SiO5 calculated using the program TWQ (Berman, 1988, 1990, 1991). Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.

8 The Phase Rule in Metamorphic Systems
2)  Equilibrium has not been attained The phase rule applies only to systems at equilibrium, and there could be any number of minerals coexisting if equilibrium is not attained Graywackes typically contain fragments of a number of different rocks in a matrix of grains derived from a rapidly eroding source Dozens of minerals may be present, but they are not in equilibrium with one another A number of igneous and metamorphic rocks also have retrograde reactions that begin, but do not run to completion

9 The Phase Rule in Metamorphic Systems
3)  We didn’t choose the # of components correctly Some guidelines for an appropriate choice of C Begin with a 1-component system, such as CaAl2Si2O8 (anorthite), there are 3 common types of major/minor components that we can add a)  Components that generate a new phase Adding a component such as CaMgSi2O6 (diopside), results in an additional phase: in the binary Di-An system diopside coexists with anorthite below the solidus May sound silly, but a very real problem, and is difficult to evaluate It is not sufficient to chemically analyze a rock and assume that each constituent is a component. This would lead to dozens of components, including trace elements In order to manageably treat equilibrium mineral assemblages, it is helpful to limit the number of components and phases with which we must deal to those that are essential – this is the key: what is essential? At end: If a component goes only into a single phase (for example, P2O5  apatite, or TiO2  ilmenite) we can effectively ignore both the component and the phase because the phase rule is unaffected by reducing f and C by the same amount As a result, many minor elements, such as P2O5 and TiO2, are often ignored along with the corresponding accessory phase so as to simplify the application of the phase rule Otherwise components that generate additional phases must be considered

10 The Phase Rule in Metamorphic Systems
3)  We didn’t choose the # of components correctly b) Components that substitute for other components Adding a component such as NaAlSi3O8 (albite) to the 1-C anorthite system would dissolve in the anorthite structure, resulting in a single solid-solution mineral (plagioclase) below the solidus Fe and Mn commonly substitute for Mg Al may substitute for Si Na may substitute for K These components may just dissolve in phases (particularly if they are not present in great quantities) We may decide to combine components that substitute for one another into a single, mixed component such as (Fe + Mg + Mn) As a rule we don’t include trace elements in the phase rule, as they behave as substitutes only and are not present in sufficient quantity that they affect the number and nature of the mineral phases present Although there is no need to ignore or combine major components like this (providing that we understand the nature of the mineral solutions in our rocks) we shall see that it is a handy device to limit the number of components when we wish to construct diagrams to display chemical-mineralogical data

11 The Phase Rule in Metamorphic Systems
3)  We didn’t choose the # of components correctly c) “Perfectly mobile” components Mobile components are either a freely mobile fluid component or a component that dissolves readily in a fluid phase and can be transported easily The chemical activity of such components is commonly controlled by factors external to the local rock system They are commonly ignored in deriving C for metamorphic systems

12 The Phase Rule in Metamorphic Systems
Consider the very simple metamorphic system, MgO-H2O Possible natural phases in this system are periclase (MgO), aqueous fluid (H2O), and brucite (Mg(OH)2) How we deal with H2O depends upon whether water is perfectly mobile or not A reaction can occur between the potential phases in this system: MgO + H2O  Mg(OH)2 Per + Fluid = Bru retrograde reaction as written (occurs as the rock cools and hydrates)

13 Cool to the temperature of the reaction curve, periclase reacts with water to form brucite: MgO + H2O  Mg(OH)2 1) Suppose H2O is mobile and we ignore it as a component If we don’t treat it as a component, we cannot treat it as a phase either (fluid is usually gone by the time we look at the rock anyway) Begin with periclase above the reaction temperature phi = 1 (water doesn’t count) and C = 1 (MgO), so the mineralogical phase rule holds As we cool to the reaction curve, periclase reacts with water to form brucite: MgO + H2O  Mg(OH)2 Reaction: periclase coexists with brucite: phi = C + 1 = 2 F = 1 (#2 reason to violate the mineralogical phase rule) To leave the curve, all the periclase must be consumed by the reaction, and brucite is the solitary remaining phase phi = 1 and C = 1 again so phi = C “Perfectly mobile” means that H2O can be added as it is needed, and any excess will leave as it is produced Water = external reservoir, available enough to enable magnesium to exist as either periclase or brucite, depending on which is stable under the given hydrous P-T conditions Imagine the diagram to be a field area, and the reaction an isograd For any common sample phi = 1 (ignore H2O and don’t see fluid in the rock), and phi = C (mineralogical phase rule) Only an uncommon sample, collected exactly on the isograd, will have phi > C Figure P-T diagram for the reaction brucite = periclase + water. From Winter (2010). An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.

14 2) If water is not perfectly mobile, we must treat the system differently
Water is not a freely permeating phase: it may become a limiting factor in the reaction Imagine cooling periclase until it intersects the reaction curve At this point we stress an obvious, but commonly overlooked, aspect of chemical reactions: A reaction involving more than one reactant will proceed until any one of the reactants is consumed As the system cools, the reaction will proceed until either periclase or water is consumed When one runs out, the other has nothing more to react with Periclase can react only with the quantity of water that can diffuse into the system If water is not perfectly mobile, and is limited in quantity, it will be consumed first Once the water is gone, the excess periclase remains stable into the brucite stability field Reactions such as this do represent absolute stability of brucite- the only reactant in the prograde reaction (brucite  periclase + water) Reaction is not the absolute stability of periclase, or of water: either can be stable across the boundary if the other reactant is absent Of course this is true for any reaction involving multiple phases Figure P-T diagram for the reaction brucite = periclase + water. From Winter (2010). An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.

15 We can now conclude that periclase can be stable anywhere on the whole diagram, if water is present in insufficient quantities to permit the reaction to brucite to go to completion In water-deficient systems we are likely to find brucite and periclase coexisting under the conditions to the left of the univariant curve We find that phi = 2 over a range of metamorphic grades, and thus over a significant portion of the field outcrops as well C = 2 as well now since water is not perfectly mobile and it becomes a limiting factor and must therefore be counted phi = C still holds At any point (other than on the univariant curve itself) we would expect to find two phases, not one phi = brucite + periclase below the reaction curve (if water is limited), or periclase + water above the curve To the right of the curve we appear to have only periclase, but if we count water as a component, we must also include it as a phase: the fluid phase Figure P-T diagram for the reaction brucite = periclase + water. From Winter (2010). An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.

16 The Phase Rule in Metamorphic Systems
How do you know which way is correct? The rocks should tell you Phase rule = interpretive tool, not predictive If only see low-f assemblages (e.g. Per or Bru in the MgO-H2O system) ® some components may be mobile If many phases in an area it is unlikely that all is right on univariant curve, and may require the number of components to include otherwise mobile phases, such as H2O or CO2, in order to apply the phase rule correctly

17 Chemographic Diagrams
Chemographics refers to the graphical representation of the chemistry of mineral assemblages A simple example: the plagioclase system as a linear C = 2 plot: = 100 An/(An+Ab) Binary Ab-An solid solution along a line Metamorphic components typically expressed as molar quantities, and not as weight % Any intermediate composition is simply plotted an appropriate distance along the line An50 would plot in the center, An25 would plot ¼ of the way from Ab to An, etc.

18 Chemographic Diagrams
3-C mineral compositions are plotted on a triangular chemographic diagram as shown in Fig x, y, z, xz, xyz, and yz2 Suppose you had a small area of a metamorphic terrane in which the rocks correspond to a hypothetical 3-component system with variable proportions of the components x-y-z The rocks in the area are found to contain 6 minerals with the fixed compositions x, y, z, xz, xyz, and yz2

19 Suppose that the rocks in our area have the following 5 assemblages:
x - xy - x2z xy - xyz - x2z xy - xyz - y xyz - z - x2z y - z - xyz When applying such a chemographic diagram to the study of mineral assemblages in a particular area, phases that coexist at equilibrium in a rock are connected by tie-lines Figure Hypothetical three-component chemographic compatibility diagram illustrating the positions of various stable minerals. Minerals that coexist compatibly under the range of P-T conditions specific to the diagram are connected by tie-lines. After Best (1982) Igneous and Metamorphic Petrology. W. H. Freeman.

20 Note that this subdivides the chemographic diagram into 5 sub-triangles, labeled (A)-(E)
x - xy - x2z xy - xyz - x2z xy - xyz - y xyz - z - x2z y - z - xyz A diagram like this is a compatibility diagram, a type of phase diagram commonly employed by metamorphic petrologists Any point within the diagram represents a specific bulk rock composition The diagram determines the corresponding mineral assemblage that develops at equilibrium For example, a point within the sub-triangle (E), the corresponding mineral assemblage corresponds to the corners = y - z - xyz Any rock with a bulk composition plotting within triangle (E) will develop that same mineral assemblage

21 Common point corresponds to 3 phases, thus f = C
Figure Hypothetical three-component chemographic compatibility diagram illustrating the positions of various stable minerals. Minerals that coexist compatibly under the range of P-T conditions specific to the diagram are connected by tie-lines. After Best (1982) Igneous and Metamorphic Petrology. W. H. Freeman. The lever principle works too for proportions of the minerals Compatibility diagrams explain clearly why rocks, even though equilibrated at the same metamorphic grade, often develop different mineral assemblages A shift in bulk composition, even if only slightly, can vary the mineral assemblage If move from (E) to (D) a rock will contain the mineral x2z (along with z and xyz) and no longer the mineral y that occurred in (E) Only rocks with bulk compositions corresponding to sub-triangle (A), for example, develop the mineral x, because they are rich in the x component Suppose some mineral, such as mineral xyz, is an important index mineral (e.g. garnet). The diagram can readily explain why all pelitic rocks need not contain garnet, even if they are within the garnet zone Any rock with a bulk composition in sub-triangle (A) will be garnet-free The diagram also demonstrates why some mineral pairs do not coexist. Mineral x cannot coexist at equilibrium with y, xyz, or z under the conditions represented, because tie-lines separate these pairs Likewise mineral z cannot coexist with x, xz, or xyz

22 What happens if you pick a composition that falls directly on a tie-line, such as point (f)?
Figure Hypothetical three-component chemographic compatibility diagram illustrating the positions of various stable minerals. Minerals that coexist compatibly under the range of P-T conditions specific to the diagram are connected by tie-lines. After Best (1982) Igneous and Metamorphic Petrology. W. H. Freeman. In this case the mineral assemblage consists of xyz and z only (the ends of the tie-line), since by adding these two phases together in the proper proportion you can produce the bulk composition (f) In such a situation phi = 2 …but so does C

23 In the unlikely event that the bulk composition equals that of a single mineral, such as xyz, then f = 1, but C = 1 as well “compositionally degenerate”

24 Chemographic Diagrams
Valid compatibility diagram must be referenced to a specific range of P-T conditions, such as a zone in some metamorphic terrane, because the stability of the minerals and their groupings vary as P and T vary Previous diagram refers to a P-T range in which the fictitious minerals x, y, z, xy, xyz, and x2z are all stable and occur in the groups shown At different grades the diagrams change Other minerals become stable Different arrangements of the same minerals (different tie-lines connect different coexisting phases)

25 A diagram in which some minerals exhibit solid solution
Figure Hypothetical three-component chemographic compatibility diagram illustrating the positions of various stable minerals, many of which exhibit solid solution. After Best (1982) Igneous and Metamorphic Petrology. W. H. Freeman. Minerals x(y,z) and x2(y,z) show limited solid solution of components y and z on one type of lattice site. Mineral x(y,z) allows more y in the lattice than does mineral x2(y,z) Minerals (xyz)ss and zss (the subscript denotes solid solution) show limited solid solution of all three components Click Suppose a bulk rock composition is in the shaded field of the mineral (xyz)ss phi = 1, but the system is not degenerate Due to the variable nature of the composition of the phase, C must still equal 3 and the phase rule tells us that F = C - phi + 2 = 4 Thus P, T, and any 2 of the 3 components in the phase are independently variable The shaded area of mineral (xyz)ss is thus an area (compositionally divariant), and our single-phase rock can have any composition within the shaded solid solution limits phi < C in this case because of the solid solution and compositional variance

26 If Xbulk on a tie-line Figure Hypothetical three-component chemographic compatibility diagram illustrating the positions of various stable minerals, many of which exhibit solid solution. After Best (1982) Igneous and Metamorphic Petrology. W. H. Freeman. As in the previous example, there are two coexisting phases (xyz)ss and zss The composition of the two minerals that correspond to bulk rock composition (f) are indicated by the two shaded dots at the ends of the tie-line through (f) phi = 2 and C is still 3, so F = = 3 Since P and T are independently variable, that means the composition of each phase is univariant, and must vary along the lines where the bundles of tie- lines end Although the composition of (xyz)ss can vary anywhere in the shaded area, the composition of (xyz)ss that coexists with Zss is constrained to the edge of the area facing z A degree of freedom is thus lost as a phase is gained Likewise the composition of zss that coexists with (xyz)ss is constrained to a portion of the edge of the shaded zss area

27 Xbulk in 3-phase triangles F = 2 (P & T) so Xmin fixed
Figure Hypothetical three-component chemographic compatibility diagram illustrating the positions of various stable minerals, many of which exhibit solid solution. After Best (1982) Igneous and Metamorphic Petrology. W. H. Freeman. In such situations phi = 3, and C = 3, as predicted by the mineralogical phase rule Since F = 2 and corresponds to P and T the phase rule tells us that all of the compositional variables for each phase are fixed

28 Chemographic Diagrams for Metamorphic Rocks
Most common natural rocks contain the major elements: SiO2, Al2O3, K2O, CaO, Na2O, FeO, MgO, MnO and H2O such that C = 9 Three components is the maximum number that we can easily deal with in two dimensions What is the “right” choice of components? Some simplifying methods: All 9 is clearly too complex. Must simplify if we want to display the system in a convenient graphical way Four components requires 3-D tetrahedra, and we lose even a semi- quantitative sense of depth in the diagram More than four components is much too complex to be useful

29 1) Simply “ignore” components Trace elements
Elements that enter only a single phase (we can drop both the component and the phase without violating the phase rule) Perfectly mobile components Some minerals are stable over a large range of metamorphic conditions, and thus are not useful as indicators of metamorphic grade Even if present in significant quantities, they may be eliminated from consideration without impairing the analysis, particularly if they are the only phase containing a particular component This is often the case for albite, K-feldspar, magnetite, ilmenite, sphene, and apatite

30 2) Combine components Components that substitute for one another in a solid solution: (Fe + Mg) 3) Limit the types of rocks to be shown Only deal with a sub-set of rock types for which a simplified system works 4) Use projections I’ll explain this shortly

31 The phase rule and compatibility diagrams are rigorously correct when applied to complete systems
A triangular diagram thus applies rigorously only to true (but rare) 3-component systems If drop components and phases, combine components, or project from phases, we face the same dilemma we faced using simplified systems in Chapters 6 and 7 Gain by being able to graphically display the simplified system, and many aspects of the system’s behavior become apparent Lose a rigorous correlation between the behavior of the simplified system and reality The advantages of being able to see what is going on by adopting an appropriate and easily visualized system, generally outweighs the loss due to simplification When we simplify, however, the diagrams may at times exhibit some “inconsistencies” If we choose a good simplified system, inconsistencies will be minimal When inconsistencies occur, careful analysis using a lesser degree of simplification usually reveals the reason for it

32 The ACF Diagram Illustrate metamorphic mineral assemblages in mafic rocks on a simplified 3-C triangular diagram Concentrate only on the minerals that appeared or disappeared during metamorphism, thus acting as indicators of metamorphic grade Eskola (1915) proposed the ACF diagram as a way to illustrate metamorphic mineral assemblages in mafic rocks on a simplified 3-C triangular diagram He wanted to concentrate only on the minerals that appeared or disappeared during metamorphism, thus acting as indicators of metamorphic grade

33 Figure 24. 4. After Ehlers and Blatt (1982). Petrology. Freeman
Figure After Ehlers and Blatt (1982). Petrology. Freeman. And Miyashiro (1994) Metamorphic Petrology. Oxford. Fig illustrates the positions of several common metamorphic minerals on the ACF diagram. Note: this diagram is presented only to show you where a number of important phases plot. It is not specific to a P-T range and therefore is not a true compatibility diagram, and has no petrological significance

34 The ACF Diagram The three pseudo-components are all calculated on an atomic basis: A = Al2O3 + Fe2O3 - Na2O - K2O C = CaO P2O5 F = FeO + MgO + MnO To calculate A for a mineral, we combine the atomic proportions Al and Fe3+ in the mineral formula, and then subtract Na and K If you begin with an ideal mineral formula Fe3+ is rare, except in a few minerals In real mafic minerals, Fe is probably in both valence states, but generally is dominated by Fe2+ Microprobe analyses cannot distinguish Fe oxidation states, but you can often estimate the relative proportions of ferric and ferrous iron by calculating the charge balance

35 The ACF Diagram A = Al2O3 + Fe2O3 - Na2O - K2O Why the subtraction?
Na and K in the average mafic rock are typically combined with Al to produce Kfs and Albite In the ACF diagram, we are interested only in the other K- bearing metamorphic minerals, and thus only in the amount of Al2O3 that occurs in excess of that combined with Na2O and K2O (in albite and K-feldspar) Because the ratio of Al2O3 to Na2O or K2O in feldspars is 1:1, we subtract from Al2O3 an amount equivalent to Na2O and K2O in the same 1:1 ratio Although it appears as though we are subtracting sodium and potassium from aluminum, we are really subtracting from Al an amount of Al that is equivalent to the amounts of Na + K, and then dealing with the Al that is left over As we’ll see in Section , this subtraction is more appropriately considered as a projection from K-feldspar and albite, thereby eliminating K2O and Na2O as components Combing Al and Fe3+ is not very rigorous, and projecting from K-feldspar is only justified when this phase is present, which is rare in mafic rocks The amounts of Fe3+ and K2O are usually minor, however, and this is seldom a major problem Projecting from Ab is more reliable, because this component is commonly present in plagioclase

36 The ACF Diagram C = CaO - 3.3 P2O5 F = FeO + MgO + MnO
C is formulated in a similar way to A All the P2O5 in most rocks is combined with CaO (in the ratio 1:3.3) to create apatite Apatite is an ubiquitous accessory mineral, and in the ACF diagram we are interested only in the CaO that exists in excess of that captured by P2O5 to create apatite Thus we subtract from CaO an amount equal to 3.3 times the amount of P2O5 to eliminate P2O5 as a component and apatite as a phase without altering CaO in the rest of the system F is a combined pseudo-component based on the common exchangeability of Fe, Mg, and Mn in solid solution on the octahedral sites of most mafic minerals If they behave in such a similar fashion, it makes some sense to treat them as a single component

37 The ACF Diagram By creating these three pseudo-components, Eskola reduced the number of components in mafic rocks from 8 to 3 Water is omitted under the assumption that it is perfectly mobile Note that SiO2 is simply ignored We shall see that this is equivalent to projecting from quartz In order for a projected phase diagram to be truly valid, the phase from which it is projected must be present in the mineral assemblages represented Thus, to be valid, the ACF diagram must have both quartz, alkali feldspar, and plagioclase present It may work anyway when these phases are lacking, but the result may also violate the mineralogical phase rule on occasion

38 The ACF Diagram An example: Anorthite CaAl2Si2O8
A = = 1, C = = 1, and F = 0 Provisional values sum to 2, so we can normalize to 1.0 by multiplying each value by ½, resulting in A = 0.5 C = 0.5 F = 0 Anorthite thus plots half way between A and C on the side of the ACF triangle, as shown in Fig. 24.4 Where does Ab plot? Plagioclase?

39 Figure 24. 4. After Ehlers and Blatt (1982). Petrology. Freeman
Figure After Ehlers and Blatt (1982). Petrology. Freeman. And Miyashiro (1994) Metamorphic Petrology. Oxford. Where does K-feldspar plot?? For KAlSi3O8 A = 0.5 + 0 - 0.5 = 0, C = 0, and F = 0 Kspar doesn’t plot on the ACF diagram If you try this for albite you will find that it doesn’t plot either The ACF formula eliminates Na and K from the diagram (and thus these phases) without altering the remaining Al after the removal All Ca-bearing plagioclase feldspars, regardless of the K and Na content, plot at the anorthite point

40 A typical ACF compatibility diagram, referring to a specific range of P and T (the kyanite zone in the Scottish Highlands) Plot all phases and connect coexisting ones with tie-lines The composition of most mafic rocks fall in the hornblende-plagioclase field or the hornblende- plagioclase-garnet triangle, and thus most metabasaltic rocks occur as amphibolites or garnet amphibolites in this zone More aluminous rocks develop kyanite and/or muscovite and not hornblende More calcic rocks lose Ca-free garnet, and contain diopside, grossularite, or even calcite (if CO2) We again see how the diagram allows us to interpret the relationship between the chemical composition of a rock and the equilibrium mineral assemblage Figure After Turner (1981). Metamorphic Petrology. McGraw Hill.

41 The AKF Diagram In the AKF diagram, the pseudo-components are:
Because pelitic sediments are high in Al2O3 and K2O, and low in CaO, Eskola proposed a different diagram that included K2O to depict the mineral assemblages that develop in them In the AKF diagram, the pseudo-components are: A = Al2O3 + Fe2O3 - Na2O - K2O - CaO K = K2O F = FeO + MgO + MnO Note that CaO is now subtracted from Al2O3 in calculating A This eliminates CaO and plagioclase from the diagram by projecting We are now interested only in the Al that occurs in excess of that combined with K, Na, and Ca to make any feldspar As a result, no feldspar, including anorthite or any intermediate plagioclase, plots on the AKF diagram Only one atom of Ca is subtracted from two of Al because the Al:Ca ratio in anorthite is 1:2.

42 Figure 24.6. After Ehlers and Blatt (1982). Petrology. Freeman.
Another figurative AKF to show where minerals plot

43 AKF compatibility diagram (Eskola, 1915) illustrating paragenesis of pelitic hornfelses, Orijärvi region Finland Figure After Eskola (1915) and Turner (1981) Metamorphic Petrology. McGraw Hill. Al-poor rocks contain biotite, and may contain an amphibole if sufficiently rich in Mg and Fe, or microcline if not Rocks richer in Al contain andalusite and cordierite

44 Three of the most common minerals in metapelites: andalusite, muscovite, and microcline, all plot as distinct points in the AKF diagram And & Ms plot as the same point in the ACF diagram, and Micr doesn’t plot at all, so the ACF diagram is much less useful for pelitic rocks (rich in K and Al)

45 Projections in Chemographic Diagrams
When we explore the methods of chemographic projection we will discover: Why we ignored SiO2 in the ACF and AKF diagrams What that subtraction was all about in calculating A and C It will also help you to better understand the AFM diagram and some of the shortcomings of projected metamorphic phase diagrams

46 Projection from Apical Phases
Example- the ternary system: CaO-MgO-SiO2 (“CMS”) Straightforward: C = CaO, M = MgO, and S = SiO2… none of that fancy subtracting business! Let’s plot the following minerals: Fo - Mg2SiO4 Per - MgO En - MgSiO3 Qtz - SiO2 Di - CaMgSi2O6 Cc - CaCO3 A naturally simple system that occurs in metamorphosed siliceous dolomites (+ CO2) and some ultramafic rocks (with some Fe and other relatively minor components)

47 Projection from Apical Phases
Fo - Mg2SiO4 Per - MgO En - MgSiO3 Qtz - SiO2 Di - CaMgSi2O6 Cc - CaCO3 For the purpose of illustration, let’s now project the composition of Di from the Ca apex of the triangle to the Mg-Si side. To do this we draw a line from C through the Di point to the M-S side of the triangle (labeled Di’)

48 The line intersects the M-S the side at a point equivalent to 33% MgO
The line intersects the M-S the side at a point equivalent to 33% MgO 67% SiO2 Note that any point on the dashed line from C through Di to the M-S side has a constant ratio of Mg:Si = 1:2 Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.

49 Projection from Apical Phases
Fo - Mg2SiO4 Per - MgO En - MgSiO3 Qtz - SiO2 Di - CaMgSi2O6 Cc - CaCO3 Pseudo-binary Mg-Si diagram in which Di is projected to a 33% Mg - 66% Si + Cal By projecting we have reduced the ternary system to a pseudo-binary one, thus making it simpler The geometric projection of diopside from CaO is mathematically equivalent to ignoring CaO in the formulas The ratio of MgO to SiO2 in Di is 1:2 or 33:67 We could calculate and plot Di in the pseudo-binary system directly without going through the complicated procedure of plotting the 3-C system, drawing the (dashed) projection line, and then redrawing the result MgO SiO 2 Fo En Di' Q Per

50 Projection from Apical Phases
Could project Di from SiO2 and get C = 0.5, M = 0.5 Di’ plots at the midpoint of the Ca-Mg side We get the same result mathematically by ignoring CaO in the diopside formula Note that periclase, forsterite, and enstatite all stack up on the same point in this projection, impeding its usefulness + Qtz MgO CaO Di' Cal Per, Fo, En

51 Projection from Apical Phases
MgO SiO 2 Fo En Di' Q Per In accordance with the mineralogical phase rule (f = C) get any of the following 2-phase mineral assemblages in our 2-component system: Per + Fo Fo + En En + Di Di + Q Suppose next that we were to treat the pseudo-binary as a compatibility diagram for some rocks belonging to the CMS system End: Because this diagram is projected from CaO, we must include calcite in each of the above assemblages if we were to consider the full 3-C (and hence 3-f) system from which we projected It appears that quartz cannot coexist with enstatite, nor can diopside coexist with forsterite Imagine our surprise as we check back with our field notes and discover that enstatite and quartz commonly coexist, as do diopside and forsterite

52 Projection from Apical Phases
What’s wrong? Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. In order to find out, we need to go back to the more rigorously correct 3-C system (hypothetical) In the complete diagram we see that our projection from CaO collapsed several important tie-lines, and that calcite does not coexist with each of the assemblages listed above To project from some point assumes that a phase plotting at that point coexists with the projected phases Thus the projected diagram assumes calcite is present In 3-C we have Di + En + Fo (coexisting diopside and forsterite) and Di + En + Q (coexisting enstatite and quartz), each of which is a calcite-free 3-phase assemblage By referring back to the “real” system, we understand why the projected system gave us some spurious results The apparent incompatibility of En + Q and Di + Fo is an artifact of the projection, which assumes that calcite is the third phase in each projected assemblage! Because calcite cannot coexist with forsterite or enstatite, these phases should not be projected from calcite at the CaO apex to the MgO-SiO2 system Projected from Calcite + Cal MgO SiO 2 Fo En Di' Q Per

53 Projection from Apical Phases
What’s wrong? Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. Better to have projected from Diopside + Di MgO SiO 2 Per Fo En Q

54 Projection from Apical Phases
ACF and AKF diagrams eliminate SiO2 by projecting from quartz Math is easy: projecting from an apex component is like ignoring the component in formulas The shortcoming is that these projections compress the true relationships as a dimension is lost To be properly applicable, quartz must therefore be present in every mineral assemblage on the ACF and AKF diagrams (or SiO2 must at least be present to the point of saturation) Why create projected pseudo- systems when they lead to such inconsistencies? 3-C -> 2-C does not provide much of a benefit, since both are easily illustrated in 2D diagrams When > 3 components, the systems are no longer so easily visualized Reducing the components from 8 or 9 to three provides a motivation to combine components As seen in the ACF and AKF diagrams, the ability to visualize the geometric relationships is often an advantage that outweighs the problems

55 Projection from Apical Phases
Two compounds plot within the ABCQ compositional tetrahedron, x (formula ABCQ) y (formula A2B2CQ) Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. To better understand how this applies to projecting from SiO2 to triangular ACF or AKF diagrams, consider the fictitious 4-component system ABCQ illustrated in Fig

56 Projection from Apical Phases
x = ABCQ y = A2B2CQ Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. If minerals coexisting within a metamorphic zone are connected by tie-lines, the tetrahedron becomes subdivided into triangular-sided sub-polyhedra: AYQX, BYQX, etc A bulk rock composition that falls within any one polyhedral volume will consist of the 4- phase equilibrium assemblage represented by the 4 corners of the sub-polyhedron

57 Projection from Apical Phases
x = ABCQ y = A2B2CQ Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. Now visualize a projection from the Q apex to the ABC plane The point X is in the center of the tetrahedron and the point Y is within the ABQX volume The projection process works the same way as in the ternary example above Projecting from Q projects points X and Y to the ABC base as points X' and Y’ Along any single projection line, such as Q-X-X' or Q-Y-Y', the ratio of A:B:C is constant We can mathematically perform the projection by simply ignoring Q in the formula of the projected phase and normalize A+B+C to 100 For X the result is A:B:C = 1:1:1 = 33:33:33 so that X plots as X' in the center of the projected pseudo- ternary diagram Y also plots as Y' with A:B:C = 2:2:1 = 40:40:20, resulting in the pseudo-ternary ABC system illustrated in Fig

58 Projection from Apical Phases
x plots as x' since A:B:C = 1:1:1 = 33:33:33 y plots as y' since A:B:C = 2:2:1 = 40:40:20 x = ABCQ y = A2B2CQ Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. The resulting ABC projection is simplified with respect to the quaternary system, and we need no longer attempt to estimate the depth of points X and Y But the projection suffers from the same pitfalls that we saw above

59 Projection from Apical Phases
If we remember our projection point (q), we conclude from this diagram that the following assemblages are possible: (q)-b-x-c (q)-a-x-y (q)-b-x-y (q)-a-b-y (q)-a-x-c The assemblage a-b-c appears to be impossible

60 Projection from Apical Phases
Figure Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall. Note the sub-polyhedron a-b-c-x in 4-C Phases a, b, and c are truly on the base and x (and y) are above it The diagram in Fig is not truly ternary, but a projection, which must be remembered! The pseudo-ternary triangle is really part of a quaternary system Because it is projected from q, it should be clearly referenced to coexisting q!

61 Projection from Apical Phases
Although A, B, and C are separated by several tie-lines in Fig , and appear to be incompatible together, they can coexist, but with x and without q! We’re not just dropping Q when we create the projection in Fig The sub-triangles in the figure are only rigorously valid groupings when phase Q is included as a part of each mineral assemblage Conscientious workers include the message “+Qtz” in the corner of most diagrams, such as ACF, AKF, etc. that are projected from quartz Multiple projections are possible, each successive one reducing the number of remaining components The ACF diagram is projected from SiO2 and then from alkali feldspar In such cases all projection phases should be noted in the diagrams

62 J.B. Thompson’s A(K)FM Diagram
An alternative to the AKF diagram for metamorphosed pelitic rocks Although the AKF is useful in this capacity, J.B. Thompson (1957) noted that Fe and Mg do not partition themselves equally between the various mafic minerals in most rocks Pelites are sensitive to T & P, and undergo many mineralogical changes as grade increases Good to have a comprehensive diagram to display these changes

63 J.B. Thompson’s A(K)FM Diagram
Figure Partitioning of Mg/Fe in minerals in ultramafic rocks, Bergell aureole, Italy After Trommsdorff and Evans (1972). A J Sci 272, Talc is always richer in Mg than coexisting diopside, diopside more so than tremolite, etc All but anthophyllite are more Mg-rich than olivine The same type of partitioning occurs in virtually all mafic minerals, including those in pelitic rocks Garnet, staurolite, and chloritiod have low Mg/Fe ratios, whereas cordierite has a higher ratio Thompson concluded that it would be advantageous in many situations to account for this unequal distribution, since it plays a role in mineral stability -> AFM diagram, in which Fe and Mg are no longer combined as a single component He neglected minor components in pelitic rocks and considered water to be perfectly mobile He eliminated SiO2 by projecting from quartz Since quartz is nearly always present in metapelites, this requirement of the projection is generally met

64 J.B. Thompson’s A(K)FM Diagram
A = Al2O3 K = K2O F = FeO M = MgO Four principal components thus remain (A = Al2O3, K = K2O, F = FeO, M = MgO) In order to avoid dealing with a three-dimensional tetrahedron, Thompson projected the phases in the system to the AFM face, thereby eliminating K2O He recognized that projecting from K2O would not work, since no phase corresponds to this apical point, and projections would cross important tie-lines

65 J.B. Thompson’s A(K)FM Diagram
Project from a phase that is present in the mineral assemblages to be studied Since muscovite is the most widespread K-rich phase in metapelites, he decided to project from muscovite (Mu) to the AFM base as shown Projecting from muscovite can lead to a strange looking AFM projections Note that Mu is still rather K-poor, and only mineral phases in the volume A-F-M-Mu in Fig will be projected to points within the AFM face of the AKFM tetrahedron Biotite is outside this volume, and projecting it from Mu causes it to plot as a band (of variable Fe/Mg) outside the AFM triangle Figure AKFM Projection from Mu. After Thompson (1957). Am. Min. 22,

66 J.B. Thompson’s A(K)FM Diagram
At high grades muscovite dehydrates to K-feldspar as the common high-K phase Then the AFM diagram should be projected from K-feldspar When projected from Kfs, biotite projects within the F-M base of the AFM triangle Muscovite may be absent in some pelitic rocks, particularly at higher grades when it dehydrates, giving way to K-feldspar as the common high-K phase When this is the case, the AFM diagram should be projected from K-feldspar if the mineral assemblages of the diagram are to have significance When projected from Kfs, biotite projects within the F-M base of the AFM triangle Figure AKFM Projection from Kfs. After Thompson (1957). Am. Min. 22,

67 J.B. Thompson’s A(K)FM Diagram
A = Al2O3 - 3K2O (if projected from Ms) = Al2O3 - K2O (if projected from Kfs) F = FeO M = MgO We avoid having to create diagrams like this, and attempting 3-D projections, by calculating the projected components of the AFM diagram mathematically, using the following formulas: We can then plot any composition as a point directly on the AFM diagram When A, F, and M are all positive, plotting a point is straightforward When A is negative, however, as with biotite, the process requires an extra step:

68 J.B. Thompson’s A(K)FM Diagram
Biotite (from Ms): KMg2FeSi3AlO10(OH)2 A = 0.5 - 3 (0.5) =  - 1 F = 1 M = 2 To normalize we multiply each by 1.0/(2 + 1 - 1) = 1.0/2 = 0.5 Thus A = -0.5 F = 0.5 M = 1 To plot the point, we extend a line from A at a constant M/F ratio. Since M/(F+M) = 0.66 We next extend a vertical line from A a distance equal to half the distance from A to the F-M base, but we extend this distance beyond the base, because A is negative Thus it extends to A/(A+F+M) = and not 0.5 Where a horizontal line at this value of A intersects the first line of constant F:M is the location of our biotite on the AFM diagram The broad biotite field in Fig is due to Al-Al substitution for (Fe-Mg)-Si in biotite K-feldspar, when projected from muscovite, projects away from the AFM diagram In order to include Kfs in the AFM diagram it is assigned a position at negative infinity

69 J.B. Thompson’s A(K)FM Diagram
Figure AFM Projection from Ms for mineral assemblages developed in metapelitic rocks in the lower sillimanite zone, New Hampshire After Thompson (1957). Am. Min. 22, Mg-enrichment typically in the order: cordierite > chlorite > biotite > staurolite > garnet Due to extensive Mg-Fe solid solution in biotite and garnet, much of the area is dominated by 2-phase fields with tie- lines (really 4-phase when we include the Qtz and Mu projection phases) Although we can easily plot ideal mineral formulas on the ACF and AKF diagrams, in order for a real mafic phase to be plotted on an AFM diagram we must know Mg/(Fe+Mg), which can only be determined by chemical analysis of the minerals, generally performed using the electron microprobe If analyses are unavailable, we can approximate the correct positions on the basis of typical relative Mg/(Fe+Mg), based on our knowledge of numerous analyses of these minerals available in the literature. From these we know that Mg-enrichment occurs typically in the order: cordierite > chlorite > biotite > staurolite > garnet

70 Choosing the Appropriate Chemographic Diagram
Example, suppose we have a series of pelitic rocks in an area. The pelitic system consists of the 9 principal components: SiO2, Al2O3, FeO, MgO, MnO, CaO, Na2O, K2O, and H2O How do we lump those 9 components to get a meaningful and useful diagram? The ACF, AKF, and AFM diagrams are the most commonly used chemographic diagrams, but several others have been employed for rocks of different compositions The choice of components and projection point depends on the system and mineral assemblages under study

71 Choosing the Appropriate Chemographic Diagram
Each simplifying step makes the resulting system easier to visualize, but may overlook some aspect of the rocks in question MnO is commonly lumped with FeO + MgO, or ignored, as it usually occurs in low concentrations and enters solid solutions along with FeO and MgO In metapelites Na2O is usually significant only in plagioclase, so we may often ignore it, or project from albite As a rule, H2O is sufficiently mobile to be ignored as well We are left with SiO2, Al2O3, K2O, CaO, FeO, and MgO Which combination and diagram is best?

72 Choosing the Appropriate Chemographic Diagram
Common high-grade mineral assemblage: Sil-St-Mu-Bt-Qtz-Plag Figure AFM Projection from Ms for mineral assemblages developed in metapelitic rocks in the lower sillimanite zone, New Hampshire After Thompson (1957). Am. Min. 22, A bulk rock composition corresponding to this assemblage is marked with an x in the AFM diagram shown previously

73 Choosing the Appropriate Chemographic Diagram
Sil-St-Mu-Bt-Qtz-Plag Figure After Ehlers and Blatt (1982). Petrology. Freeman. In the AFM diagram Sil, St, and Bt plot (with Mu + Qtz as associated phases, required by the projection) In the ACF diagram Ky, St, and Plag are all that plot In both the ACF and AFM diagrams C = P = 3, so we have “valid” divariant assemblages as far as the mineralogical phase rule is concerned The ACF diagram is less useful for pelites, because plagioclase is the only Ca-bearing phase, and is stable over a wide range of conditions Plag is thus not much of an indicator of metamorphic grade Biotite, which is an important pelitic mineral that is often indicative of both grade and composition, does not appear Although our assemblage plots as a 3-phase sub-triangle, other pelitic assemblages plot on the A-F edge, and most of the area of the diagram is wasted The ACF diagram is more useful for mafic rocks

74 Choosing the Appropriate Chemographic Diagram
Sil-St-Mu-Bt-Qtz-Plag We don’t have equilibrium There is a reaction taking place (F = 1) We haven’t chosen our components correctly and we do not really have 3 components in terms of AKF The AKF diagram contains 4 phases and the Sil-Bt and Mu-St tie-lines cross Crossing tie-lines can be a troublesome problem, because we cannot choose a single sub-triangle for our rock composition x Remember, there are three possible explanations for this seeming violation of the mineralogical phase rule now that f > C We don’t have equilibrium There is a reaction taking place (F = 1) We haven’t chosen our components correctly and we do not really have 3 components in terms of AKF Figure After Ehlers and Blatt (1982). Petrology. Freeman.

75 Choosing the Appropriate Chemographic Diagram
Sil-St-Mu-Bt-Qtz-Plag That Mu+Ky+Bt+St is a common assemblage argues against first two possibilities If we compare the AKF and AFM diagrams we see Fe/Mg partitioning between Bt & St is moderately strong, and should not be ignored Adding more Mg to the biotite than the staurolite effectively lifts Bt higher than St from the pure Fe end of the AKF base As a result, the 4 phases appearing in the AKF diagram are not really coplanar in a more comprehensive AKFM 4-component diagram Thus the Sil-Bt and Mu-St tie-lines do not intersect in space, so that Sil+Bt is not compositionally equivalent to Mu+St We conclude that the AFM diagram is the best choice for the assemblage in question It may not be the best for other assemblages, however, and the AFM and AKF diagrams are often used in combination to depict the relationship between rock composition, mineral assemblage, and metamorphic grade in metamorphosed pelitic rock sequences Figure After Ehlers and Blatt (1982). Petrology. Freeman.

76 Choosing the Appropriate Chemographic Diagram
Myriad chemographic diagrams have been proposed to analyze paragenetic relationships in various metamorphic rock types Most are triangular: the maximum number that can be represented easily and accurately in two dimensions Some natural systems may conform to a simple 3- component system, and the resulting metamorphic phase diagram is rigorous in terms of the mineral assemblages that develop Other diagrams are simplified by combining components or projecting These diagrams are not as rigorous, but the ability to compare chemical and mineralogical variations usually outweighs the disadvantages, particularly when the method of simplification is clearly understood Anomalies, such as phi appearing to be greater than C and crossed tie-lines may thus be resolved by referring to the more comprehensive system

77 Choosing the Appropriate Chemographic Diagram
Variations in metamorphic mineral assemblages result from: 1) Differences in bulk chemistry 2) differences in intensive variables, such as T, P, PH2O, etc (metamorphic grade) A good chemographic diagram permits easy visualization of the first situation The second can be determined by a balanced reaction in which one rock’s mineral assemblage contains the reactants and another the products These differences can often be visualized by comparing separate chemographic diagrams, one for each grade


Download ppt "Chapter 24. Stable Mineral Assemblages in Metamorphic Rocks"

Similar presentations


Ads by Google