12.07.2015 Views

an empirically derived kinetic model for albitization of detrital ...

an empirically derived kinetic model for albitization of detrital ...

an empirically derived kinetic model for albitization of detrital ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase315Fig. 1. Equilibrium boundary <strong>of</strong> <strong>albitization</strong> according to reaction 2 (see text), after Boles (1982). Theerror in the equilibrium const<strong>an</strong>t calculated from free energy values given by Robie <strong>an</strong>d Waldbaum (1968) is1 pH unit. We assume quartz provides silica. The plot shows that reservoir fluids are in the albite stabilityfield <strong>an</strong>d yet <strong>albitization</strong> was not observed.words, the <strong>albitization</strong> zone is thermodynamically constrained by the pore fluidchemistry in basins. The reaction onset, however, is <strong>kinetic</strong>ally controlled <strong>an</strong>d proceedsat <strong>an</strong> appreciable rate after a critical activation energy is reached, presumably attemperatures higher th<strong>an</strong> 70°C.relev<strong>an</strong>t <strong>kinetic</strong> experimentation <strong>an</strong>d <strong>model</strong>ingThere is a long tradition <strong>of</strong> dissolution <strong>an</strong>d precipitation experiments <strong>an</strong>d<strong>model</strong>ing involving feldspars; Helgeson <strong>an</strong>d others (1984), Hellm<strong>an</strong>n (1994), <strong>an</strong>dBlum <strong>an</strong>d Stillings (1995) give excellent reviews. Most experiments on albite precipitationhave been per<strong>for</strong>med at green schist <strong>an</strong>d lower amphibolite P-T-X conditions(250-500°C <strong>an</strong>d up to 1 kbar), generally in 0.1 M NaCl <strong>an</strong>d 0.05 M Na 2 SiO 3 solutions(Matthews, 1980; Moody <strong>an</strong>d others, 1985). Alkali metal halides, such as NaCl <strong>an</strong>dCaCl, have a dramatic effect on reaction <strong>kinetic</strong>s (Winkler <strong>an</strong>d Luttge, 1999), <strong>an</strong>d itmay be reasonable to consider that this is the case <strong>of</strong> subsurface environments.However, these temperature conditions are not observed in sedimentary basins <strong>an</strong>dthese experiments do not truly reproduce the nature <strong>of</strong> a pseudomorphic replacement(Merino <strong>an</strong>d others, 1993).Previous <strong>model</strong>ing <strong>of</strong> plagioclase <strong>albitization</strong> includes Baccar <strong>an</strong>d others (1993)<strong>an</strong>d Wilson <strong>an</strong>d others (2000). The <strong>for</strong>mer study focused on the minimum time <strong>of</strong><strong>albitization</strong> from 60 to 150° C in a closed system. It allowed a critical saturation be<strong>for</strong>enucleation <strong>an</strong>d growth, <strong>an</strong>d assumed <strong>an</strong> initial fluid volume, fluid composition, <strong>an</strong>dgrain size distribution. The rate const<strong>an</strong>ts were <strong>derived</strong> using specific dissolution <strong>an</strong>dprecipitation <strong>kinetic</strong> const<strong>an</strong>ts <strong>an</strong>d thermodynamic equilibrium const<strong>an</strong>ts. Baccar <strong>an</strong>dothers (1993) concluded that the <strong>albitization</strong> rate decreases with the extent <strong>of</strong> thereaction (which increases with temperature) <strong>an</strong>d that plagioclase is not stable at highPco 2 conditions. These results give insight into dissolution-precipitation processes,


316 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>such as the effect <strong>of</strong> Pco 2 , but not into replacement mech<strong>an</strong>isms. The preservation <strong>of</strong>plagioclase internal details such as twin pl<strong>an</strong>es <strong>an</strong>d cleavages, requires, among stillm<strong>an</strong>y unknown parameters, at least: a) plagioclase to dissolve <strong>an</strong>d albite to grow atequal <strong>an</strong>d slow volumetric rates (Merino <strong>an</strong>d others, 1993), b) a stress-controlledcoupled dissolution/precipitation mech<strong>an</strong>ism (Merino <strong>an</strong>d Dewers, 1998), c) a structurallyself-org<strong>an</strong>ized fluid film at the mineral interface (O’Neil <strong>an</strong>d Taylor, 1967;Reyh<strong>an</strong>i <strong>an</strong>d others, 1999), <strong>an</strong>d d) differences between parent <strong>an</strong>d daughter crystalsolubility (Putnis, 2002).The latter study, (Wilson <strong>an</strong>d others, 2000), considered <strong>albitization</strong> as a processindependent <strong>of</strong> fluid composition. The <strong>model</strong> used <strong>an</strong> activation energy <strong>of</strong> 65 kJ/molbased on plagioclase dissolution experiments at low temperatures (Blum <strong>an</strong>d Stillings,1995) <strong>an</strong>d fit the reaction order <strong>an</strong>d the frequency factor to a single compositionaldata (Boles <strong>an</strong>d Ramseyer, 1988). This empirical approach was useful to simulate thechemical evolution <strong>of</strong> the pore fluids in the S<strong>an</strong> Joaquin Basin. It also proved to besuccessful by yielding results consistent with geologic observations. The <strong>model</strong>, however,lacked a chemical affinity term, validation in other settings, <strong>an</strong>d was alsoindependent <strong>of</strong> a surface area variable. After reviewing these two studies, we concludedthat <strong>an</strong> <strong>albitization</strong> <strong>kinetic</strong> <strong>model</strong> could be better <strong>for</strong>mulated by including a surfacearea variable, a saturation index term, <strong>an</strong>d by testing it widely against data from severalbasins.hypothesisWe postulate that the <strong>albitization</strong> rate is a function <strong>of</strong> time, temperature, fluidcomposition, <strong>an</strong>d surface area. Our rate equation follows a <strong>kinetic</strong> <strong>for</strong>mulationconsistent with the Rate Law <strong>an</strong>d the Arrhenius equation (Lasaga, 1984). Rate lawsqu<strong>an</strong>tify the dependence <strong>of</strong> reaction rates on the concentration <strong>of</strong> the startingmaterial. They take the general <strong>for</strong>m:dCdt kC n (3)where dC/dt is proportional to the ch<strong>an</strong>ge in concentration <strong>of</strong> C raised to the power <strong>of</strong>n. The Arrhenius equation qu<strong>an</strong>tifies ch<strong>an</strong>ges in the rate const<strong>an</strong>t k with temperature<strong>an</strong>d has the <strong>for</strong>m:k Ae Ea/Rt (4)where A is the frequency factor, E a the activation energy, R the universal gas const<strong>an</strong>t,<strong>an</strong>d T temperature. In our work, the <strong>kinetic</strong> parameters (E a <strong>an</strong>d A) present within thereaction rate k are fitted using solely molar compositional trends with depth, which areimplicit functions <strong>of</strong> time <strong>an</strong>d temperature. In reality, plagioclase composition, porefluid pH, <strong>an</strong>d temperature are the main controlling factors <strong>of</strong> the rate const<strong>an</strong>t.However, different plagioclase compositions occur together in samples with identicalpore fluid <strong>an</strong>d temperature history, strongly suggesting that the chemical composition<strong>of</strong> the plagioclase has a stronger effect on the rate const<strong>an</strong>t th<strong>an</strong> pore fluid (Ramseyer<strong>an</strong>d others, 1992). Furthermore, based on the following observations:(1) pore water from marine basins have a limited initial variation in pH values <strong>an</strong>dchemical composition (Fisher <strong>an</strong>d Boles, 1990); <strong>an</strong>d(2) oil field brines are almost always in equilibrium or saturated with respect toalbite <strong>an</strong>d unsaturated with respect to <strong>an</strong>orthite (Helgeson <strong>an</strong>d others, 1993),We assumed that (1) <strong>albitization</strong> is a dissolution reaction, which should have <strong>an</strong> E abetween 40 <strong>an</strong>d 80 kJ/mol (Lasaga, 1984), <strong>an</strong>d (2) during <strong>albitization</strong> the sum <strong>of</strong> the<strong>an</strong>orthite, orthoclase, <strong>an</strong>d albite mole fraction is one at all times.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase317The Saturation Index HypothesisThe dissolution rate <strong>of</strong> a given plagioclase composition depends on the rateconst<strong>an</strong>t k, <strong>an</strong>d the saturation index . In our <strong>model</strong> the rate const<strong>an</strong>t k is the fittedparameter <strong>an</strong>d we will deal with it in more detail later. The saturation index , atafixed temperature <strong>an</strong>d water composition, is commonly defined as: IAP(5)K eqwhere IAP is the ionic activity product <strong>of</strong> the solution, <strong>an</strong>d K eq the equilibriumconst<strong>an</strong>t. From the log 10 positive numbers represent supersaturation <strong>an</strong>d negativeundersaturation (<strong>for</strong> example Bethke, 1996). Based on field observations that suggestthat <strong>an</strong>orthite dissolution is the <strong>albitization</strong> rate-limiting step we assume that reaction 2is adequate to represent <strong>albitization</strong> throughout geological time in sedimentary basins.Again, reaction 2 may be written as:2quartz 0.5water H Na <strong>an</strong>orthite albite 0.5dickite Ca 2 (6)in which the IAP is defined by:IAP 1/2a Ca a dickite a albitea <strong>an</strong>orthite a Na a H a 1/2 H2O a quartzwhere the a’s are the activities <strong>of</strong> both, the solids <strong>an</strong>d aqueous species. In this work, thesolid phases were considered pure with unit activity, except <strong>an</strong>orthite. The <strong>an</strong>orthitecontent <strong>of</strong> plagioclase being albitized is 0.35 mole fraction, <strong>an</strong>d its activity is 0.45(Saxena <strong>an</strong>d Ribbe, 1972). The equilibrium const<strong>an</strong>t K eq is calculated from the Gibbsfree energy <strong>of</strong> the reaction. The free energy ch<strong>an</strong>ge at <strong>an</strong>y T <strong>an</strong>d P <strong>an</strong>d solutioncomposition Q, in its simplest <strong>for</strong>m, is given by: r G T,P r H° 298 298T r Cpdt T rS° 298 298T r CpT2 (7)dt 1P r Vdp RTlnQ (8)where r G T,P is the Gibbs free energy r H° 298 enthalpy, r S° 298 entropy, r Cp heatcapacity, <strong>an</strong>d r V the molar volume ch<strong>an</strong>ge <strong>for</strong> the reaction, all in calories/mole (table1). The subscript 298 refers to the reference temperature 298.15 Kelvin. At equilibrium,Q becomes K eq , <strong>an</strong>d the free energy ch<strong>an</strong>ge <strong>of</strong> the reaction is zero so that: TTTlnK eq rH° 298 r Cpdt 298 rCprS° 298T298dt P r RT ln0.45 (9)Vdp1As shown above, the equilibrium const<strong>an</strong>t <strong>for</strong> <strong>an</strong>y reaction is ultimately a function <strong>of</strong> P,T, fluid composition, <strong>an</strong>d solids composition. To calculate K eq <strong>for</strong> reaction 2 we usedthermodynamic data from two different, but consistent, sources. Free energies <strong>of</strong><strong>for</strong>mation, enthalpies <strong>of</strong> <strong>for</strong>mation, <strong>an</strong>d third law entropies <strong>for</strong> all solid phases wereobtained from Robie <strong>an</strong>d Waldbaum (1968). We used the database reported by Robie<strong>an</strong>d Waldbaum (1968) instead <strong>of</strong> the more recent <strong>an</strong>d st<strong>an</strong>dard database SUPCRT92(Johnson <strong>an</strong>d others, 1992) <strong>for</strong> the following reasons:


318 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Table 1Formation water compositions associated with partial or complete <strong>albitization</strong>. Data compiled from several sourcesConcentrations in mg/liter*Reported as SiTGCB Gulf Coast BasinSJB S<strong>an</strong> Joaquin BasinNI No In<strong>for</strong>mationReferences 1) Boles (1982); 2) Fisher <strong>an</strong>d Boles (1990).Temperature in degrees CelsiusDepth in meters


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase319a) Consistency with previous studies. Boles (1982) determined the <strong>albitization</strong>stability field <strong>an</strong>d errors in the free energy <strong>of</strong> <strong>for</strong>mation <strong>an</strong>d equilibrium const<strong>an</strong>t<strong>of</strong> reaction 2 using the Robie <strong>an</strong>d Waldbaum (1968) database.b) Contrary to Robie <strong>an</strong>d Waldbaum’s database, SUPCRT92 lacks st<strong>an</strong>dard enthalpy<strong>an</strong>d Gibbs free energy data <strong>for</strong> Dickite, which is part <strong>of</strong> the products in thereaction under study.d) Robie <strong>an</strong>d Waldbaum’s database contains thermodynamic data <strong>for</strong> Dmisteinbergite,<strong>an</strong> hexagonal <strong>an</strong>orthite polymorph used, as we will explain later, as a proxyto evaluate the influence <strong>of</strong> the atomic arr<strong>an</strong>gement <strong>an</strong>d crystalline structure onthe tr<strong>an</strong>s<strong>for</strong>mation <strong>kinetic</strong>s.On the other h<strong>an</strong>d, the st<strong>an</strong>dard Gibbs free energy, enthalpy <strong>an</strong>d entropy <strong>of</strong><strong>for</strong>mation <strong>of</strong> <strong>an</strong>orthite, low albite, <strong>an</strong>d quartz present in Robie <strong>an</strong>d Waldbaum (1968)differ by less th<strong>an</strong> 1.4 percent from those present in SUPCRT92, so, in spite <strong>of</strong> theother differences, we felt com<strong>for</strong>table using heat capacity data from SUPCRT92 assummarized by Anderson <strong>an</strong>d Crear (1993).In order to keep our <strong>model</strong> simple <strong>an</strong>d ease the calculation <strong>of</strong> the equilibriumconst<strong>an</strong>t we made several simplifications. First, justified by the narrow r<strong>an</strong>ge <strong>of</strong>temperatures being considered here, we assumed that the heat capacity <strong>for</strong> ions wasconst<strong>an</strong>t. We also assumed that our <strong>model</strong> is pressure independent, <strong>an</strong>d the net solidvolume ch<strong>an</strong>ge is close to zero (Boles, 1982) thereby assuming P 1 bar in equation(9) <strong>an</strong>d neglecting the pressure correction term. Consequently, lnK eq is left as afunction <strong>of</strong> temperature <strong>an</strong>d c<strong>an</strong> be represented by the following polynomial equation:log 10 K eq p 1 T 3 p 2 T 2 p 3 T p 4 (10)where p 1 1.7513 10 7 , p 2 2.7925 10 4 , p 3 0.1685, <strong>an</strong>d p 4 37.993 (fig.2). The error in the equilibrium const<strong>an</strong>t at 298.15 K, calculated from st<strong>an</strong>dard errorestimation <strong>for</strong>mulas is 0.98. It is likely that this error may exp<strong>an</strong>d at highertemperatures. The thermodynamic values used are summarized in table 2. As we willdiscuss further in the text, we pl<strong>an</strong> to qualitatively examine the effect <strong>of</strong> the <strong>an</strong>orthitecrystalline structure on <strong>albitization</strong>, in other words, on reaction 2. Thus, figure 2 alsoillustrates the log 10 K eq <strong>for</strong> reaction 2 assuming that <strong>an</strong> hexagonal <strong>an</strong>orthite (<strong>an</strong><strong>an</strong>orthite polymorph) <strong>detrital</strong> grain is being albitized, in which case we use differentst<strong>an</strong>dard free energy <strong>an</strong>d entropy values (table 2). Heat capacity data <strong>for</strong> hexagonal<strong>an</strong>orthite are, to our knowledge, not available, so we retained the same heat capacityequations. The equilibrium const<strong>an</strong>t <strong>for</strong> reaction 2 assuming hexagonal <strong>an</strong>orthite c<strong>an</strong>also be fitted to equation (10) with different coefficients: p 1 1.4412 10 7 , p 2 2.2995 10 4 , p 3 0.13938, <strong>an</strong>d p 4 30.421. The error in log 10 K eq is larger th<strong>an</strong> 0.98, but no further calculations were per<strong>for</strong>med.Speciation <strong>an</strong>d ion activities were, on the other h<strong>an</strong>d, calculated withSOLMINEQ.88 (Kharaka <strong>an</strong>d Barnes, 1973; Perkins <strong>an</strong>d others, 1990). We usedreservoir fluid compositions from the S<strong>an</strong> Joaquin <strong>an</strong>d Gulf Coast Basins, at depthintervals where <strong>albitization</strong> is present <strong>an</strong>d thermodynamically active (see our table 1;data from Boles, 1982; Fisher <strong>an</strong>d Boles, 1990). Based on the water chemistry data weassumed that the oil field waters are in equilibrium with quartz, so that the activity <strong>of</strong>the solid quartz activity is one. This assumption is justified by the presence <strong>of</strong>authigenic quartz in intervals above the <strong>albitization</strong> zone, also by fluid chemistry data(Wilson <strong>an</strong>d others, 2000). The SOLMINEQ.88 code is only partially consistent withthe origins <strong>of</strong> the thermodynamic database presented in table 2. However, the<strong>an</strong>alyzed Na <strong>an</strong>d Ca 2 concentrations in the reservoir waters differ by less th<strong>an</strong> 10percent from those concentrations calculated by the code, clearly indicating that the


320 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Fig. 2. Plot showing the equilibrium const<strong>an</strong>t <strong>for</strong> reaction 2 assuming pure <strong>an</strong>orthite <strong>an</strong>d its hexagonalpolymorph (see text). The log 10 K is calculated with equation (9) as a function <strong>of</strong> temperature. Thermodynamicdata is from Robie <strong>an</strong>d Waldbaum (1968) <strong>an</strong>d Johnson <strong>an</strong>d others (1992). The error in log 10 K(<strong>an</strong>orthite) at 298.15 K <strong>an</strong>d 1 bar, calculated from st<strong>an</strong>dard error estimation <strong>for</strong>mulas <strong>an</strong>d from thermochemicalerrors in the <strong>for</strong>mer database is 0.98. It is likely that this error may exp<strong>an</strong>d at higher temperatures.complexation <strong>of</strong> these ions in minerals is negligible. Consequently, <strong>an</strong>y thermodynamiccalculation <strong>of</strong> Na <strong>an</strong>d Ca 2 activities would be consistent with the thermodynamicdatabase used in our paper. The resulting activities, ionic activity products, <strong>an</strong>dequilibrium const<strong>an</strong>ts <strong>for</strong> a variety <strong>of</strong> fluid compositions <strong>an</strong>d temperatures used in our<strong>model</strong> are shown in table 3.The <strong>for</strong>mer table shows the variation <strong>of</strong> the saturation index from S<strong>an</strong> Joaquin<strong>an</strong>d Gulf Coast Basin fluids that are currently at temperatures between 100 <strong>an</strong>d 130° C,<strong>an</strong>d where <strong>albitization</strong> is actively occurring (Boles, 1982). However, considering thewaters’ ages are considerably different, in general several millions <strong>of</strong> years <strong>of</strong> difference,<strong>an</strong>d that they have gone through different time-temperature paths, it could beconsidered that the variation <strong>of</strong> is relatively small. Thus, given 1) the narrow r<strong>an</strong>ge <strong>of</strong>


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase321Table 2Thermodynamic values used to calculate the equilibrium const<strong>an</strong>t <strong>for</strong> reaction 2*Delta <strong>of</strong> the reaction assuming <strong>an</strong>orthite**Delta <strong>of</strong> the reaction assuming hexagonal <strong>an</strong>orthiteG° St<strong>an</strong>dard Gibbs Free Energy <strong>of</strong> Formation (Calories/mole)S° St<strong>an</strong>dard Entropy “Third Law” <strong>of</strong> phase (Calories/K mole)H° St<strong>an</strong>dard Enthalpy <strong>of</strong> Formation (Calories/mole)a Calories/moleb Calories/mole K 2c Calories/mole K 5The sum considers the stoichiometry <strong>of</strong> the reactionHeat capacity <strong>of</strong> pure phases <strong>an</strong>d thermodynamic data <strong>for</strong> aqueous ions from SUPCRT92 <strong>an</strong>dsummarized by Anderson <strong>an</strong>d Crear (1993)compositions <strong>of</strong> albitized <strong>detrital</strong> plagioclase grains, 2) the relatively small variations<strong>for</strong> waters <strong>of</strong> considerably different ages (millions <strong>of</strong> years) <strong>an</strong>d different T-t paths, <strong>an</strong>d3) the occurrence <strong>of</strong> different plagioclase compositions in samples with identical porefluid <strong>an</strong>d temperature history, we hypothesize that a weighted average integration <strong>of</strong> asaturation index as a function <strong>of</strong> temperature, may be useful to simplify the variations<strong>of</strong> water chemistry <strong>of</strong> the studied basins.All our assumptions may seem, a priori, very simplistic <strong>an</strong>d far from realistic but they areconstrained by field data <strong>an</strong>d petrographic observations, <strong>an</strong>d they are not unreasonable. First,normally all replacement reactions are considered isovolumetric (Merino <strong>an</strong>d others,1993; Nahon <strong>an</strong>d Merino, 1997; Merino <strong>an</strong>d Dewers, 1998). Second, it might bejustified that during the pseudomorphic replacement, the coupled dissolutionprecipitation<strong>of</strong> plagioclase <strong>an</strong>d albite, establishes locally a small-scale steady state or atleast a “quasi” steady state composition <strong>of</strong> the pore fluid, leading to a local saturationthat could be assumed to be about const<strong>an</strong>t over long time intervals. This idea wouldbe particularly justified if the Al 3 <strong>an</strong>d Ca 2 leave the local system at a const<strong>an</strong>t rate<strong>an</strong>d fluid tr<strong>an</strong>sport would not be too fast, as in the case in deep crustal settings awayfrom fractures <strong>an</strong>d faults, where permeabilities could be in the r<strong>an</strong>ge <strong>of</strong> tenths <strong>of</strong>milidarcies <strong>an</strong>d the average flow velocity is, at times, less th<strong>an</strong> 1 cm/Ma (Garven, 1995).


322 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Table 3Input data <strong>for</strong> saturation index calculation. Speciation per<strong>for</strong>med with SOLMINEQ.88IAP Ionic activity product Keq Equilibrium const<strong>an</strong>t


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase323general <strong>kinetic</strong> <strong>model</strong>s <strong>an</strong>d equationsLasaga (1984, 1995, <strong>an</strong>d 1998) explains all theoretical <strong>an</strong>d fundamental approachesto describe dissolution <strong>an</strong>d precipitation rates <strong>of</strong> minerals. The general<strong>kinetic</strong> <strong>for</strong>mulation <strong>for</strong> mineral/fluids interactions is equation (11):Rate S m Ae Ea/RT na OH n g I aii f r G (11)iwhere the Rate is a function <strong>of</strong> the reactive surface area S m (cm 2 ), the frequency factorA (1/s), the apparent overall activation energy E a (kJ/mol) the temperature T (K), theuniversal gas const<strong>an</strong>t R (J/K mol), the OH nactivity raised to the power <strong>of</strong> na OH(dimensionless), the ionic strength <strong>of</strong> the solution g(I )(I has units <strong>of</strong> molality timesthe ionic charge squared ), the activity products a n i <strong>of</strong> the rate-determining specie i(dimensionless) raised to the power <strong>of</strong> n, <strong>an</strong>d a general function <strong>of</strong> the free energy <strong>of</strong>the reaction f( r G) that c<strong>an</strong> take m<strong>an</strong>y <strong>for</strong>ms, where r G has units <strong>of</strong> J/mol. Anotherwidely used empirical <strong>kinetic</strong> <strong>model</strong>, written as <strong>an</strong> adaptation <strong>of</strong> equation (11) byAagaard <strong>an</strong>d Helgeson (1982) is:Rate S m k m a i m 1 (12)where S m is the interfacial area (cm 2 ) between the mineral <strong>an</strong>d the aqueous solution,k m is the rate const<strong>an</strong>t (1/s), a i n is the activity <strong>of</strong> the rate-determining aqueous species iin solution. On the other h<strong>an</strong>d, the <strong>kinetic</strong>s <strong>of</strong> pseudomorphic replacement have been<strong>for</strong>mulated with two simult<strong>an</strong>eous reactions that occur at equal volumetric rates,driven by a <strong>for</strong>ce <strong>of</strong> crystallization generated by supersaturation (Maliva <strong>an</strong>d Seiver,1988; Dewers <strong>an</strong>d Ortoleva, 1990; Merino <strong>an</strong>d others, 1993; Nahon <strong>an</strong>d Merino, 1997).In this case, the sum <strong>of</strong> the rates equals zero as in equation (13)R a R b 0 (13)where R i is the linear dissolution rate <strong>of</strong> mineral i in cm/s represented by equation(14):R i k i 1 (14)where k i (cm/s) <strong>an</strong>d i have the usual me<strong>an</strong>ings. Equations (11) through (14) dependon the aqueous ionic species present during the reaction <strong>an</strong>d are founded on awell-developed physical-chemical theoretical basis, <strong>an</strong>d only laboratory-based measurements,with good control <strong>of</strong> fluid composition (in either closed or open systems) c<strong>an</strong>be most easily used <strong>for</strong> a rigorous interpretation <strong>of</strong> <strong>kinetic</strong> rate laws <strong>an</strong>d reactionmech<strong>an</strong>isms. However, the prediction <strong>an</strong>d <strong>model</strong>ing <strong>of</strong> the fluid saturation path insedimentary basins is far from attainable. The variation <strong>an</strong>d evolution <strong>of</strong> chemical <strong>an</strong>dionic species through geologic time is even more difficult to predict or <strong>model</strong> withprecision. Furthermore, invalid extrapolations <strong>of</strong> experimental <strong>kinetic</strong> parameters togeological settings <strong>an</strong>d our poor underst<strong>an</strong>ding <strong>of</strong> reaction mech<strong>an</strong>isms (<strong>an</strong>d <strong>of</strong> m<strong>an</strong>yother variables) limit the applicability <strong>of</strong> rigorous <strong>an</strong>d experimental approaches t<strong>of</strong>ield examples.Based on the reasons explained previously, most empirical <strong>model</strong>s <strong>of</strong> diageneticreactions take a different <strong>an</strong>d simple approach; they reduce <strong>an</strong>d ignore a vast amount<strong>of</strong> variables, specifically fluid compositions by avoiding or leaving const<strong>an</strong>t the saturationindex or chemical affinity parameter in the rate expression. The <strong>model</strong>s lack aphysical-chemical basis (Oelkers <strong>an</strong>d others, 2000), however, they c<strong>an</strong> successfullypredict dissolution <strong>an</strong>d precipitation over long time scales, which is the ultimate goal<strong>of</strong> these studies.


324 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>The smectite to illite conversion during shale diagenesis provides one example <strong>of</strong><strong>empirically</strong> <strong>derived</strong> <strong>kinetic</strong>s. Several <strong>model</strong>s <strong>for</strong> this particular reaction have beenproposed <strong>an</strong>d evaluated (Elliot <strong>an</strong>d Matis<strong>of</strong>f, 1996). The most recent <strong>model</strong> (Hu<strong>an</strong>g<strong>an</strong>d others, 1993) computes the rate with: dSdt kK S 2 (15)where dS/dt is the ch<strong>an</strong>ge <strong>of</strong> the smectite fraction with respect to time t, k is the rateconst<strong>an</strong>t, [K ] is a fixed <strong>an</strong>d const<strong>an</strong>t potassium concentration (in ppm), <strong>an</strong>d S is thedimensionless fraction <strong>of</strong> smectite in the illite/smectite raised to the power <strong>of</strong> 2.Another example is quartz overgrowth precipitation. Commonly, <strong>model</strong>s <strong>for</strong>quartz precipitation couple diffusion, <strong>kinetic</strong>, <strong>an</strong>d thermodynamic aspects <strong>of</strong> thereaction (Oelkers, 1996; Oelkers <strong>an</strong>d others, 2000). In contrast, a simpler mathematic/<strong>kinetic</strong> <strong>model</strong> based on time-temperature <strong>an</strong>d surface area (Walderhaug, 1994 <strong>an</strong>d1996) continues to be successfully tested in basins around the world (<strong>for</strong> exampleAwwiller <strong>an</strong>d Summa, 1997; L<strong>an</strong>der <strong>an</strong>d Walderhaug, 1999; Perez <strong>an</strong>d others, 1999a,1999b). The quartz precipitation <strong>model</strong> uses a logarithmic rate function dependent ontemperature (eq 16):Rate a10 b Tt (16)where a (moles/cm 2 ) <strong>an</strong>d b (1/°C) are pre-exponential <strong>an</strong>d exponential const<strong>an</strong>ts <strong>an</strong>dT(t) is temperature expressed as a function <strong>of</strong> time. The precipitated quartz volume Vq(cm 3 ) is calculated as a sum <strong>of</strong> integrals within sequences <strong>of</strong> time steps i (eq 17):Vq M n t i1 a A b10i1it Tt idt(17)where M is quartz molar volume (g/mole), is quartz density (g/cm 3 ), A i is surfacearea (cm 2 ), <strong>an</strong>d t is time (sec). As we explained previously, these empirical <strong>model</strong>sdepend mainly on the variations <strong>of</strong> the heating rate through geologic time, obtainedthrough basin <strong>model</strong>ing (Siever, 1983), <strong>an</strong>d lack <strong>an</strong> affinity term.<strong>albitization</strong> <strong>kinetic</strong> <strong>model</strong>Based on previous <strong>for</strong>mulations in the S<strong>an</strong> Joaquin Basin (Wilson <strong>an</strong>d others,2000), we <strong>model</strong>ed <strong>albitization</strong> with a <strong>kinetic</strong>/mathematical <strong>for</strong>mulation expressed inequation (18) by: d An Sdt m kAn o 2 1 (18)where d[An]/dt is the ch<strong>an</strong>ge <strong>of</strong> the <strong>an</strong>orthite mole fraction with respect to time t, k isthe rate const<strong>an</strong>t in 1/cm 2 s, [An o ] 2 is the initial <strong>an</strong>orthite mole fraction squared, <strong>an</strong>d the weighted average saturation index. In equation (18) S m is the area <strong>of</strong> themineral-solution interface (Lasaga, 1984) <strong>an</strong>d is expressed in equation (19) in thefollowing <strong>for</strong>m:S m A m v (19)V mwhere, A m is the mineral surface area (cm 2 ), V m is the mineral volume assumingspherically-shaped grains (cm 3 ), <strong>an</strong>d v is the unit volume <strong>of</strong> the s<strong>an</strong>dstone, which


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase325includes porosity <strong>an</strong>d is equal to 1 cm 3 . Simplifying, the surface area interface results inequation (20):S m 6 d v (20)where d is the grain diameter (cm). In all <strong>an</strong>alyzed samples the grain size r<strong>an</strong>ges from0.012 0.001 to 0.086 0.018 cm, varying from one sample to <strong>an</strong>other, however,average grain diameter is 0.031 0.002 cm (table 4). These values are obtained bymeasuring the me<strong>an</strong> grain diameter in 40 grains per thin section, using a st<strong>an</strong>dardpolarized light petrographic microscope.The rate const<strong>an</strong>t k follows the Arrhenius equation (21):k A exp E a(21)RTtwhere A <strong>an</strong>d E a are apparent pre-exponential <strong>an</strong>d exponential parameters in 1/cm 2 s<strong>an</strong>d kJ/mole respectively, R the universal gas const<strong>an</strong>t (J/K mole), <strong>an</strong>d T(t) istemperature in K expressed as a function <strong>of</strong> time t.The initial <strong>an</strong>orthite molar fraction value [An o ] is obtained through field measurements.It is determined on <strong>detrital</strong> plagioclase grains with <strong>an</strong> ARL microprobe<strong>an</strong>alyzer, <strong>model</strong> SMX-SM equipped with three wavelength-dispersive spectrometers<strong>an</strong>d a Tracor Northern TN2000 energy dispersive system, by simult<strong>an</strong>eously <strong>an</strong>alyzing<strong>for</strong> sodium, potassium, <strong>an</strong>d calcium, aluminum, <strong>an</strong>d silicon. Each grain is <strong>an</strong>alyzed onthe rim <strong>an</strong>d in the core, <strong>an</strong>d 50 grains are usually <strong>an</strong>alyzed per thin section fromdifferent depth intervals <strong>an</strong>d temperatures (Boles, 1982; Boles <strong>an</strong>d Ramseyer, 1988;Pitmm<strong>an</strong>, 1988).Based on the reasons explained above we used a me<strong>an</strong> saturation index term ( ) obtained with a weighted average integration. The method to obtain is asfollows: a) first we calculate several numerical values <strong>of</strong> (defined in equation 5) usingwaters compositions from different reservoirs <strong>of</strong> different ages, depths <strong>an</strong>d burialtemperatures where <strong>albitization</strong> is active (table 3); b) observing a slight temperaturedependence <strong>of</strong> these values, we fit them to a function <strong>of</strong> temperature–which isimplicitly a function <strong>of</strong> time; c) we weight averaged the function with equation (22), sothat: TTdT dT (22)The integration along the (T ) function (solution to equation 22), neglecting theextreme values <strong>an</strong>d within the temperature <strong>of</strong> interest, results in a const<strong>an</strong>t log 10 2.632. This positive number suggests that pore fluid compositions are in the albite dickite stability field, which is consistent with field observations. In short, to start the<strong>albitization</strong>, pore fluids must be supersaturated with respect to albite, that is, in thealbite stability field according to reaction 2.It is well known that, in laboratory experiments, dissolution <strong>an</strong>d precipitationreactions are usually not independent <strong>of</strong> fluid composition, <strong>an</strong>d such assumptionwould only be valid if the reaction occurs far from equilibrium. Thus, a const<strong>an</strong>t maybe somewhat problematic. However, based on (1) the small r<strong>an</strong>ge <strong>of</strong> water composi-


326 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Table 4Summary <strong>of</strong> data sets used as input in our <strong>model</strong>. Data include unpublished wellnames, <strong>an</strong>d samples’ temperature, age, <strong>an</strong>d grain size.*Gulf Coast Basin data from Boles (1982), the S<strong>an</strong> Joaquin Basin data from Boles <strong>an</strong>d Ramseyer (1988) <strong>an</strong>d the Denver Basin data from Pittm<strong>an</strong> (1988)**Present day depth***Present day temperatureAb st<strong>an</strong>ds <strong>for</strong> ablite, Or <strong>for</strong> orthoclase <strong>an</strong>d An <strong>for</strong> <strong>an</strong>orthites.d. st<strong>an</strong>ds <strong>for</strong> st<strong>an</strong>dard deviation <strong>of</strong> spot <strong>an</strong>alysis per sample [100 spots <strong>for</strong> Boles (1982) <strong>an</strong>d Boles <strong>an</strong>d Ramseyer (1988) <strong>an</strong>d 20–27 spots <strong>for</strong> Pitm<strong>an</strong>n (1988)]N.I. No in<strong>for</strong>mationThe grain size units are in millimeters


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase327tions present in the S<strong>an</strong> Jaoquin <strong>an</strong>d Gulf Coast Basin where <strong>albitization</strong> is active(Boles, 1982), (2) a narrow <strong>albitization</strong> temperature window, (3) the relatively smallr<strong>an</strong>ge <strong>of</strong> albitized plagioclase compositions (fig. 2), we propose that the heating ratemay be the most import<strong>an</strong>t parameter <strong>for</strong> the actual <strong>kinetic</strong>s <strong>of</strong> the replacement.Particularly because experiments suggest that this is the case when the reactionproducts grow on the react<strong>an</strong>t (Luttge <strong>an</strong>d others, 1998), as in the case <strong>of</strong> replacementreactions.In addition, at depths greater th<strong>an</strong> 3 to 4 km, where <strong>albitization</strong> occurs, the rockpermeabilities are on the order <strong>of</strong> the tenths <strong>of</strong> milidarcies <strong>an</strong>d the rate <strong>of</strong> fluidtr<strong>an</strong>sport is very small; <strong>an</strong>d it is reasonable to assume that some minerals c<strong>an</strong> establishlocally a steady state or at least a “quasi” steady state composition <strong>of</strong> the pore fluid,leading to a local saturation that could be assumed to be about const<strong>an</strong>t over long timeintervals.Numerical SolutionThe solution <strong>of</strong> equation (18) is reached by the integration <strong>of</strong> equation (23):t mdAn Ae 6 Ea/R Ttv And o t 2 1 dt (23)0An <strong>an</strong>alytical integration <strong>of</strong> the right h<strong>an</strong>d side <strong>of</strong> equation (23) from time t 0 to t m isnot trivial due to the time dependence <strong>of</strong> the temperature, <strong>an</strong>d some proposedrational approximations (<strong>for</strong> example Burnham <strong>an</strong>d others, 1987) introduce errors athigh temperatures (Comer, 1992). Alternatively, a numerical solution c<strong>an</strong> be reachedif T(t) is divided in increments, <strong>an</strong>d a me<strong>an</strong> value T i is calculated within the (i) (i 1) time steps. In the general case, the me<strong>an</strong> temperature T i may be calculatedthrough a weighted average integration expressed in equation (24) as:t it i1Ttdt T i t it i1dt(24)If the increments are small <strong>an</strong>d considered linear, T(t) takes the <strong>for</strong>m <strong>of</strong> equation (25)nTt m i t d i (25)i1where the slope m is the temperature ch<strong>an</strong>ge rate (in °C/s), t is time, <strong>an</strong>d d is the initialtemperature (in °C), <strong>an</strong>d T i is reduced to <strong>an</strong> average between two time steps, that is(T i T i1 )/2. If a time-temperature curve, represented by the function TT(t), isdivided in small linear increments from (T i ,t i )to(T i1 ,t i1 ), the slope m i betweeneach interval would be:<strong>an</strong>d the initial temperature d i is:m i T i1 T it i1 t i(26)d i T i T i1 T it i1 t it i1 (27)


328 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Our <strong>model</strong> uses variable times steps, between 0.3 <strong>an</strong>d 0.7 million years, dependingupon the sample age <strong>an</strong>d temperature history. The <strong>model</strong> results are independent <strong>of</strong>the time step length, although intervals <strong>of</strong> 0.5 million years are preferred because theygive best results. The substitution <strong>of</strong> T(t)by T i results in equation (28):t i1d An Ae 6 Ea/R T v And o t 2 1 dt (28)iAlthough equations (23) <strong>an</strong>d (28) may seem similar, the substitution simplifies theintegration, because T is now a const<strong>an</strong>t. Finally, the integration is numericallyexpressed with equation (29):nAn i1 An o i Ae Ea/R T i6d vAn o 2 i 1 t i1 t i (29)i0where i is the time step.* During <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase in the studiedbasins, the <strong>an</strong>orthite mole fraction decreases from An 35 to An 0 , <strong>an</strong>d the albite molefraction increases simult<strong>an</strong>eously from Ab 65 to Ab 99 (table 4). By definition the sum <strong>of</strong>the mole fractions is unity, that is:An Ab Or 1 (30)where An, Ab, <strong>an</strong>d Or are the <strong>an</strong>orthite, albite, <strong>an</strong>d orthoclase mole fractions.There<strong>for</strong>e, the simult<strong>an</strong>eous gain <strong>of</strong> the albite component, as a result <strong>of</strong> <strong>albitization</strong>, iscomputed with equation (31)nAb i1 1 An o i Ae Ea/R T i6d v An o 2 i 1 t i1 t i Or o (31)i1where [Ab] is the final albite mole fraction per time step i,[Or o ] is the initial orthoclasemole fraction that could be set const<strong>an</strong>t to 0.05 0.01, or simply ignored.The only unknown parameters in equation 29 are E a <strong>an</strong>d A. When only twoparameters are being optimized, both least squares fit <strong>an</strong>d direct search techniques areefficient methods (Gallagher <strong>an</strong>d Ev<strong>an</strong>s, 1991). We approached the problem with adirect grid search, using <strong>an</strong> iteration method, which breaks the time-temperaturedomain into a series <strong>of</strong> discrete blocks, where equations 23 through 31 are calculated.We beg<strong>an</strong> our search <strong>of</strong> A by fixing E a from 60 to 70 kJ/mol following Walther <strong>an</strong>dWood (1984) <strong>an</strong>d Lasaga (1984). We decided to search with small E a <strong>an</strong>d wide Aincrements. Our decision was based on the limited E a r<strong>an</strong>ge (60 E a 80 kJ/mol)reported <strong>for</strong> a great number <strong>of</strong> dissolution reactions <strong>of</strong> alumino silicates (Lasaga,1984) <strong>an</strong>d specifically <strong>for</strong> feldspar hydrolysis reactions at neutral pH within thetemperature <strong>of</strong> interest (Hellm<strong>an</strong>, 1994; Blum <strong>an</strong>d Stillings, 1995).As input, we used (1) compositional data that revealed increasing <strong>albitization</strong> withdepth <strong>an</strong>d temperature (table 2), <strong>an</strong>d (2) time-temperature curves obtained through1-D burial-thermal <strong>model</strong>ing <strong>of</strong> the basins. We per<strong>for</strong>med the basin <strong>model</strong>ing usingGenesis tm (Zhiyong, 2003).*Note that [An o ] i 2 is a const<strong>an</strong>t per time step.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase329Fig. 3. Several time-temperature paths <strong>of</strong> albitized samples from the S<strong>an</strong> Joaquin Basin used in the<strong>kinetic</strong> <strong>model</strong>.<strong>model</strong> fit <strong>an</strong>d testsS<strong>an</strong> Joaquin Basin.—We beg<strong>an</strong> our search in the S<strong>an</strong> Joaquin Basin. Callaway(1971) provided <strong>an</strong> excellent review <strong>of</strong> its geologic framework. In summary, the basincontains more th<strong>an</strong> 7 km <strong>of</strong> Tertiary Miocene sediment, <strong>an</strong>d underwent down warpingby compaction during its burial history. In the center <strong>of</strong> the basin, where we focusedour study, the present depths represent maximum burial. For the <strong>albitization</strong> <strong>kinetic</strong>fit, we selected from the published compositional data (Boles <strong>an</strong>d Ramseyer, 1988) atotal <strong>of</strong> 11 samples from nine different wells (table 4). The selection was based on theavailability <strong>of</strong> thin sections <strong>an</strong>d well data. We constructed time-temperature histories<strong>for</strong> each sample <strong>an</strong>d determined grain sizes by averaging the diameter <strong>of</strong> 40 plagioclasegrains per sample.We extracted input parameters from the literature <strong>for</strong> basin <strong>model</strong>ing purposes,including stratigraphic in<strong>for</strong>mation from field-scale structural cross sections(Cali<strong>for</strong>nia Oil <strong>an</strong>d Gas Division, 1985), <strong>for</strong>mation ages (COSUNA, 1984 <strong>an</strong>d Wood<strong>an</strong>d Boles, 1991). The latter paper also provided <strong>for</strong>mation thickness <strong>an</strong>d subsidencerates. We assumed a const<strong>an</strong>t basal heat flow <strong>of</strong> 50 mW/m 2 , based on presentday thermal data (Lachenbruch <strong>an</strong>d Sass, 1980) <strong>an</strong>d geothermal studies in the S<strong>an</strong>Joaquin Basin (Dumitru, 1988; Wilson <strong>an</strong>d others, 2000) <strong>an</strong>d we used thermalconductivity values compiled in Genesis tm by Zhiyong (2003). It is import<strong>an</strong>t toemphasize that heat fluxes in the S<strong>an</strong> Joaquin Basin were likely lower th<strong>an</strong> 50mW/m 2 during active subduction, <strong>an</strong>d that thermal gradients have recovered sincethen (Dumitru, 1988). Figure 3 depicts the resulting burial histories. The <strong>kinetic</strong><strong>model</strong>ing search (fig. 4) indicates that the <strong>albitization</strong> trend c<strong>an</strong> be reasonablysimulated using <strong>an</strong> apparent activation energy <strong>of</strong> 68 4 kJ/mole <strong>an</strong>d apparentfrequency factors <strong>of</strong> (6.5 0.5) 10 3 1/cm 2 Ma. The error bars in the data points(fig. 4) represent the st<strong>an</strong>dard deviation, computed as the square root <strong>of</strong> thevari<strong>an</strong>ce, which is in turn calculated as the average squared deviation from the


330 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Fig. 4. Initial <strong>model</strong> fit to <strong>albitization</strong> in the S<strong>an</strong> Joaquin Basin. Measured <strong>an</strong>d calculated data areplotted as functions <strong>of</strong> temperature. The ■ symbol with vertical lines represents data from Boles <strong>an</strong>dRamseyer (1988) with st<strong>an</strong>dard deviation computed as the square root <strong>of</strong> the vari<strong>an</strong>ce. Symbols E, , ‚, <strong>an</strong>d▫ represent sequences <strong>of</strong> simulations per<strong>for</strong>med <strong>for</strong> different samples. The data fit <strong>an</strong> E a <strong>of</strong> 68 4 kJ/mol<strong>an</strong>d <strong>an</strong> A (6.5 0.5) 10 3 1/cm 2 Ma. The E a <strong>an</strong>d A errors represent the required deviation me<strong>an</strong> value toenclose the st<strong>an</strong>dard deviation <strong>of</strong> the compositional data.me<strong>an</strong>. The errors in E a <strong>an</strong>d A represents the r<strong>an</strong>ge <strong>of</strong> values in E a <strong>an</strong>d A required toenclose the variation in the compositional data.Tests ResultsOnce we calculated the set <strong>of</strong> unknown parameters in the S<strong>an</strong> Joaquin Basin, weproceeded to test them against plagioclase compositional trends from the Texas GulfCoast (Boles, 1982) <strong>an</strong>d Denver Basin <strong>of</strong> Colorado (Pittm<strong>an</strong>, 1988).Texas Gulf Coast Basin.—Galloway <strong>an</strong>d others (1994) provided a synthesis <strong>of</strong> thegeologic, structural, <strong>an</strong>d depositional history <strong>of</strong> the onshore Tertiary deposits presentin the southern part <strong>of</strong> the Texas Gulf Coast Basin. The stratigraphic column in thesouthern part <strong>of</strong> the Texas Gulf Coast consists <strong>of</strong> 4 to 6 km <strong>of</strong> sedimentary rocks, buriedin a simple down-warped fashion without major regional uplifts (Murray, 1960).Present-day depths <strong>an</strong>d temperatures in the basin represent, in m<strong>an</strong>y cases, maximumvalues (Boles, 1982). We used the <strong>kinetic</strong> <strong>model</strong> <strong>an</strong>d parameters <strong>derived</strong> from the S<strong>an</strong>Joaquin Basin fit to simulate the plagioclase compositional trend reported <strong>for</strong> theOligocene Frio Formation <strong>of</strong> south Texas (Boles, 1982). The data comprise eightsamples distributed over eight wells from six different counties. Grain sizes weredetermined by measuring the me<strong>an</strong> diameter <strong>of</strong> 50 plagioclase grains in thin sections(table 4).We constructed 1-D burial-thermal curves <strong>for</strong> each sample using Genesis tm(Zhiyong, 2003). From Dodge <strong>an</strong>d Posey’s (1981) set <strong>of</strong> subsurface structural crosssections, we obtained well locations, <strong>for</strong>mation tops, gamma-rays/spont<strong>an</strong>eous potentiallogs, bottom hole temperatures, <strong>an</strong>d biostratigraphic in<strong>for</strong>mation. We used thebiostratigraphy in<strong>for</strong>mation from the COSUNA (1984) chart <strong>an</strong>d the U.S. Department<strong>of</strong> Interior Biostratigraphic Chart <strong>for</strong> the Gulf <strong>of</strong> Mexico (OCS) Region (2002) toderive <strong>for</strong>mation ages. For the time-temperature reconstruction, the present day heat


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase331Fig. 5. Time-temperature history <strong>of</strong> albitized samples from the Gulf Coast Basin, <strong>derived</strong> frombasin-burial <strong>model</strong>ing.flux (53 mW/m 2 ) was set const<strong>an</strong>t from the middle Oligocene to the present. The heatflux was calculated by the code using <strong>for</strong>mation temperatures <strong>an</strong>d thermal conductivityvalues compiled by Zhiyong (2003). We calibrated our time-temperature historieswith temperature measurements taken along the depth <strong>of</strong> the wells (Dodge <strong>an</strong>d Posey,1981). Results are plotted in figure 5.Using the previous results, we <strong>model</strong>ed <strong>albitization</strong> <strong>an</strong>d computed the loss <strong>of</strong> the<strong>an</strong>orthite component (fig. 6A). Additionally, due to the fact that the orthoclasefraction in a <strong>detrital</strong> plagioclase remains const<strong>an</strong>t <strong>an</strong>d is usually less th<strong>an</strong> 5 percent, <strong>an</strong>albitized grain c<strong>an</strong> be conveniently studied as a binary system. In a binary system, whenthe compositional data is normalized to 100 percent, the vari<strong>an</strong>ce <strong>of</strong> each componentis the same, <strong>an</strong>d consequently, the st<strong>an</strong>dard deviations are equal. Thus, we <strong>model</strong>edthe gain <strong>of</strong> the albite component using the st<strong>an</strong>dard deviation from the <strong>an</strong>orthite molefraction (fig. 6B). Ten waters <strong>an</strong>alysis from fluid reservoirs that have gone throughdifferent time-temperature trajectories, have signific<strong>an</strong>tly different ages (millions oryears), <strong>an</strong>d where <strong>albitization</strong> is active, the waters have relatively similar saturationindex values compared to subsurface waters from the S<strong>an</strong> Joaquin Basin (table 3).Thus <strong>for</strong> <strong>model</strong>ing purposes, we used the same weight averaged log 10 2.632 valueas in the S<strong>an</strong> Joaquin Basin. The albite gain or <strong>an</strong>orthite loss during <strong>albitization</strong> inTertiary Gulf Coast Basin c<strong>an</strong> be simulated assuming reaction 2 with <strong>an</strong> apparentactivation energy <strong>of</strong> 68 4 kJ/mole <strong>an</strong>d apparent frequency factors <strong>of</strong> (6.5 0.5) 10 3 1/cm 2 Ma. Again, errors <strong>of</strong> E a <strong>an</strong>d A represent the values required to enclose thevariations in the compositional data. The error bars <strong>of</strong> the data (fig. 6B) represent thest<strong>an</strong>dard deviation calculated as the square root <strong>of</strong> the vari<strong>an</strong>ce.Denver Basin.—We tested the <strong>kinetic</strong> <strong>model</strong> against data from the CretaceousTerry s<strong>an</strong>dstone in the Spindel field, Denver Basin <strong>of</strong> Colorado. Pittm<strong>an</strong> (1988 <strong>an</strong>d1989) presented a geologic synthesis <strong>of</strong> the basin <strong>an</strong>d plagioclase <strong>albitization</strong> data. Thehistory <strong>of</strong> the basin is rather complex. The heat flux varied through time resulting in


332 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Fig. 6(A) Plot illustrating simulated <strong>an</strong>d measured <strong>an</strong>orthite mole fractions versus temperature in theGulf Coast Basin. Measured <strong>an</strong>d calculated data are plotted as a function <strong>of</strong> temperature. The ■ symbolrepresents data extracted from Boles (1982). Vertical lines represent st<strong>an</strong>dard deviation <strong>of</strong> 100 spot <strong>an</strong>alysesper sample, computed as the square root <strong>of</strong> the vari<strong>an</strong>ce. All other symbols represent sequences <strong>of</strong>simulations calculated <strong>for</strong> fine-grained s<strong>an</strong>dstones deposited between 28.6 <strong>an</strong>d 32.1 Ma (see legend). Fitcorresponds to E a 68 4 kJ/mole <strong>an</strong>d A (6.5 0.5) 10 3 1/cm 2 Ma. The errors represent the r<strong>an</strong>ge <strong>of</strong>values in E a <strong>an</strong>d A required to enclose the variation in the composition data.Fig. 6(B) Calculated <strong>an</strong>d measured albite mole fractions as a function <strong>of</strong> temperature in Gulf CoastBasin. The ■ symbol represents data from Boles (1982). The vertical lines represent st<strong>an</strong>dard deviations(square root <strong>of</strong> vari<strong>an</strong>ce) <strong>of</strong> <strong>an</strong>orthite component. Vertical lines represent st<strong>an</strong>dard deviation <strong>of</strong> 100 spot<strong>an</strong>alyses per sample.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase333Fig. 7. Thermal history <strong>of</strong> the Terry S<strong>an</strong>dstone, Spindle field, Denver Basin, after Pittm<strong>an</strong> (1988).Time-temperature pr<strong>of</strong>ile corresponds to a horizon at 1,325 meters depth <strong>an</strong>d was used as input <strong>for</strong><strong>albitization</strong> <strong>model</strong>.ch<strong>an</strong>ging thermal gradients <strong>an</strong>d heating rates, <strong>an</strong>d the sedimentary section wasuplifted 800 m during Early Paleocene <strong>an</strong>d Miocene time. Burial <strong>an</strong>d thermalhistories were calibrated against R o pr<strong>of</strong>iles, K/Ar data, <strong>an</strong>d multiple temperature logs(Pittm<strong>an</strong>, 1988).The resulting time-temperature history <strong>of</strong> the Terry s<strong>an</strong>dstone corresponds to asingle horizon at 1,325 m depth (fig. 7). Not having compositional data at specifically1,325 m depth, we chose the closest samples where data were available, correspondingto those in the well Rademacher No. 1 at 1,283.2 <strong>an</strong>d 1,286.9 m depth (see table 2, p.752, Pittm<strong>an</strong>, 1988). Albitized plagioclase in this setting was mainly reported in fine tovery fine-grained s<strong>an</strong>dstone. Hence, we <strong>model</strong>ed <strong>albitization</strong> using three surface areaassumptions, calculated from three different s<strong>an</strong>d grain sizes (1) fine grain, with d 0.0177 cm, (2) very fine grain with d 0.0088 cm, <strong>an</strong>d (3) fine to very fine grain withd 0.0125 cm.Having no in<strong>for</strong>mation regarding the oil field water chemistry or saturationindexes, we computed the reaction using the same log 10 2.632 used in the S<strong>an</strong>Joaquin <strong>an</strong>d Gulf Coast Basin. Again, results indicate that the <strong>albitization</strong> data c<strong>an</strong> beconsistently predicted with <strong>an</strong> apparent activation energy <strong>of</strong> 68 4 kJ/mole <strong>an</strong>dapparent frequency factors <strong>of</strong> (6.5 0.5) 10 3 1/cm 2 Ma (fig. 8). The same fittingscheme per<strong>for</strong>med with data from the S<strong>an</strong> Joaquin Basin was per<strong>for</strong>med independentlyusing data from the Louisi<strong>an</strong>a Gulf Coast <strong>an</strong>d Denver Basin, <strong>an</strong>d the resultsshowed no skew.<strong>model</strong> evaluationComparison with Other Diagenetic ReactionsWe compared the <strong>albitization</strong> rate const<strong>an</strong>t k with the rate const<strong>an</strong>ts <strong>of</strong> quartzprecipitation <strong>an</strong>d smectite-to-illite <strong>model</strong>s <strong>derived</strong> from field data <strong>an</strong>d burialtemperaturehistories by different authors (fig. 9). Although smectite-to-illite <strong>an</strong>d


334 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Fig. 8. Plot <strong>of</strong> comparison between calculated <strong>an</strong>d measured <strong>an</strong>orthite mole fractions in Denver Basin.Measured <strong>an</strong>d calculated data are plotted as a function <strong>of</strong> time. The ■ symbol represents two data points at1,283.2 <strong>an</strong>d 1,283.9 m depth (Pittm<strong>an</strong>, 1988), <strong>an</strong>d the error bar the st<strong>an</strong>dard deviation. Other symbolsrepresent sequences <strong>of</strong> simulations per<strong>for</strong>med <strong>for</strong> very fine-grained s<strong>an</strong>dstones (‚), fine-grained () <strong>an</strong>dvery fine to fine s<strong>an</strong>d ( ). Age <strong>of</strong> samples is 73.3 Ma. The data fit <strong>an</strong> E a 68 4 kJ/mole <strong>an</strong>d A(6.5 0.5) 10 3 1/cm 2 Ma. The errors in E a <strong>an</strong>d A represents the r<strong>an</strong>ge <strong>of</strong> values in E a <strong>an</strong>d A required to enclosethe variation in the composition data.quartz precipitation are all very different reactions, their reaction rates were <strong>derived</strong>with <strong>an</strong> empirical-numerical approach, similar to ours. Additionally, smectite-to-illite<strong>an</strong>d quartz overgrowth precipitation, together with <strong>albitization</strong>, constitute the maindiagenetic reactions in sedimentary basins: they are all strongly dependent on heatingrates <strong>an</strong>d surface area or grain size. In our <strong>model</strong>, the rate const<strong>an</strong>t k, is calculated withequation (32), identical to equation (4), as shown below:k Ae Ea/RT (32)The k is plotted in figure 9. There k r<strong>an</strong>ges from 1 10 21 1/cm 2 s at 25°C to1 10 18 1/cm 2 s at 200°C (fig. 9), assuming E a <strong>an</strong>d A to be 68 kJ/mol <strong>an</strong>d 6.5 10 31/cm 2 Ma <strong>an</strong>d making the appropriate unit conversion to 1/cm 2 s.Our rate const<strong>an</strong>t values are higher th<strong>an</strong> the quartz precipitation rate <strong>of</strong> Walderhaug(1994) <strong>an</strong>d smectite-illite tr<strong>an</strong>s<strong>for</strong>mation rates <strong>of</strong> Hu<strong>an</strong>g <strong>an</strong>d others (1993), <strong>an</strong>dlower th<strong>an</strong> Pytte <strong>an</strong>d Reynolds’ (1988) <strong>an</strong>d Velde <strong>an</strong>d Vaseur’s (1992) smectite to illiterates. We also compared our <strong>model</strong> to growth rates <strong>of</strong> albite (fig. 9). Aagaard <strong>an</strong>dothers (1990) estimated albite growth rates from albite dissolution rates (Knauss <strong>an</strong>dWolery, 1986) <strong>an</strong>d the principle <strong>of</strong> microscopic reversibility (Lasaga, 1984). Accordingto Aagaard <strong>an</strong>d others (1990), dissolution would not normally be the limiting factor inthe case <strong>of</strong> K-feldspar <strong>albitization</strong>. This conclusion is reached by comparing therelatively fast K-feldspar dissolution rate (Helgeson <strong>an</strong>d others, 1984) with the relativelyslow albite precipitation rate—omitting nucleation.Contrary to the K-feldspar case, in our <strong>model</strong> the rates <strong>of</strong> <strong>albitization</strong> (plagioclasedissolution) are slower th<strong>an</strong> the albite growth rates <strong>of</strong> all temperatures (fig. 9), which is


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase335Fig. 9. Comparison <strong>of</strong> the <strong>albitization</strong> rate calculated in this paper assuming E a 68 kJ/mole <strong>an</strong>d A 6.5 10 3 1/cm 2 Ma)–with other rates <strong>of</strong> mineral reactions as a function <strong>of</strong> temperature. For illitization <strong>of</strong>smectite (S/I) based on Pytte <strong>an</strong>d Reynolds (1988), Velde <strong>an</strong>d Vassseur (1992) <strong>an</strong>d Hu<strong>an</strong>g <strong>an</strong>d others(1993); quartz precipitation based on Walderhaug’s (1994); albite growth rate based on Aagaard <strong>an</strong>d others(1990).fairly consistent with field observations. In sedimentary basins, plagioclase dissolutionis the rate-limiting step <strong>of</strong> <strong>albitization</strong> based on the fact that low albite precipitates inporous s<strong>an</strong>dstones at temperatures lower th<strong>an</strong> the temperature onset <strong>of</strong> <strong>albitization</strong>(Ramseyer <strong>an</strong>d others, 1992).Sensitivity Studies <strong>an</strong>d Model LimitationsWe per<strong>for</strong>med several sensitivity <strong>an</strong>alyses to illustrate the <strong>model</strong>’s response tovariations in the input parameters. The common input to all simulations was <strong>an</strong>arbitrarily chosen time-temperature pr<strong>of</strong>ile, somewhat similar to the thermal history <strong>of</strong>the Gulf Coast <strong>an</strong>d S<strong>an</strong> Joaquin Basins (fig. 10). From the simulation results (figs. 11<strong>an</strong>d 12), we are able to evaluate the effects <strong>of</strong> the initial composition <strong>an</strong>d the surfacearea on <strong>albitization</strong>. Additionally, based on the Denver <strong>an</strong>d Gulf Coast Basin results, wediscuss the effect <strong>of</strong> heating rates on <strong>albitization</strong>.Effects <strong>of</strong> initial composition.—The effect <strong>of</strong> the initial <strong>an</strong>orthite [An 0 ] compositionon the rates <strong>of</strong> <strong>albitization</strong> is depicted in figure 11. From 0 to 65°C, the reaction ratesare relatively slow <strong>an</strong>d all <strong>model</strong>ed initial compositions, An 30 Ab 65 ,An 40 Ab 60 ,An 60 Ab 30<strong>an</strong>d pure An 100 , remain const<strong>an</strong>t. Between 75 <strong>an</strong>d 125°C all compositions are partiallyalbitized. Finally, at temperatures over 125°C, the different compositional trendsconverge to a single value, that is [An] 0 mole percent. It is clear that theplagioclase’s <strong>detrital</strong> composition in terms <strong>of</strong> <strong>an</strong>orthite content affects the reaction,but the effect is minimal at high temperatures. The wide st<strong>an</strong>dard deviations at lowtemperature <strong>an</strong>d small st<strong>an</strong>dard deviations at high temperature (figs. 4 <strong>an</strong>d 6) <strong>of</strong> thedata (table 1), c<strong>an</strong> both be partially explained by differences in the initial <strong>an</strong>orthitefraction.


336 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Fig. 10. Time-temperature pr<strong>of</strong>ile representing a single burial cycle, similar to pr<strong>of</strong>iles from the S<strong>an</strong>Joaquin <strong>an</strong>d Gulf Coast Basins. This curve was used as input <strong>for</strong> several sensitivity <strong>an</strong>alyses illustrated infigures 8 <strong>an</strong>d 9.Another rate determining factor to examine, but well beyond the scope <strong>of</strong> thispaper, is the influence <strong>of</strong> the ordered (that is plutonic) versus disordered (that isvolc<strong>an</strong>ic) atomic arr<strong>an</strong>gement <strong>of</strong> the <strong>albitization</strong> <strong>kinetic</strong>s. At present we know <strong>of</strong> noFig. 11. Sensitivity <strong>of</strong> <strong>albitization</strong> <strong>kinetic</strong> <strong>model</strong> to plagioclase initial composition, as a function <strong>of</strong>temperature using pr<strong>of</strong>ile in figure 7. The initial plagioclase composition is function <strong>of</strong> An. Simulations showthat initial composition persists until 100°C, slows <strong>albitization</strong> at temperatures below 125°C, <strong>an</strong>dconverges to a single composition trend over 135°C.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase337Fig. 12. Sensitivity <strong>of</strong> <strong>albitization</strong> to <strong>detrital</strong> grain size as a function <strong>of</strong> temperature. The initialplagioclase composition is common to all simulations. Simulations indicate that grain size differences affectthe degree <strong>of</strong> <strong>albitization</strong> throughout the tr<strong>an</strong>s<strong>for</strong>mation. Effects <strong>of</strong> the grain size persist until hightemperatures are reached, but it minimizes at temperatures over 160°C.thermodynamic data <strong>of</strong> ordering at const<strong>an</strong>t <strong>an</strong>orthite content in plagioclase. Thus asa proxy, we used thermodynamic data <strong>of</strong> <strong>an</strong> hexagonal polymorph named Dmisteinbergite.This polymorph is rare but has been reported on fractured surfaces in coal fromthe Chelyabinsk coal basin, Southern Ural Mountains, Russia.As mentioned the thermochemical data <strong>for</strong> hexagonal <strong>an</strong>orthite is taken fromRobie <strong>an</strong>d Waldbaum (1968). Hence, we calculated the equilibrium const<strong>an</strong>ts <strong>of</strong>reaction 2 assuming <strong>an</strong> hexagonal <strong>an</strong>orthite <strong>an</strong>d graphed it in figure 2. We alsocalculated values <strong>for</strong> reaction 2 under the same assumption <strong>an</strong>d using fluidcompositions from the S<strong>an</strong> Joaquin Basin <strong>an</strong>d Gulf Coast Basin (table 3). The resultingvalues <strong>of</strong> log 10 hexagonal suggest that if Dmisteinbergite was present in the S<strong>an</strong> JoaquinBasin, reaction 2 would proceed far enough to the right until the saturation point,either because the pore fluids have been in contact with sufficient amounts <strong>of</strong> themineral or have reacted with the mineral long enough. In short, pore fluids aresaturated with respect to hexagonal <strong>an</strong>orthite, <strong>an</strong>d we c<strong>an</strong> only speculate that the<strong>albitization</strong> <strong>of</strong> this polymorph would be faster, assuming the same rate const<strong>an</strong>t.Effect <strong>of</strong> surface area.—The mineral/fluid surface area interface is key in diagenetic<strong>model</strong>ing (Oelkers, 1996). Our <strong>model</strong> is designed such that plagioclase dissolutionrate is inversely proportional to the grain diameter. A simple <strong>an</strong>d direct inspection <strong>of</strong>equation (18) shows that d[An]/dt S n so that [An] i1 (6/d)v. Thus, given the sameinitial composition <strong>an</strong>d time-temperature path, smaller grains will albitize faster th<strong>an</strong>bigger grains. Figure 12 illustrates <strong>albitization</strong> trends plotted against temperature <strong>for</strong>several grain diameters. Below 60°C the influence <strong>of</strong> the initial grain size on <strong>albitization</strong>is small, but over 70°C the tr<strong>an</strong>s<strong>for</strong>mation rate increases as grain size decreases. Asa result plagioclase with different grain sizes may have different degrees <strong>of</strong> <strong>albitization</strong><strong>an</strong>d/or <strong>albitization</strong> paths that persist over a wide temperature r<strong>an</strong>ge. Perhaps the widedeviations <strong>of</strong> the <strong>albitization</strong> trends may be explained by the st<strong>an</strong>dard deviations ingrain size distributions (table 4). However, the dissolution <strong>of</strong> minerals, <strong>an</strong>d particularly


338 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>feldspars, is certainly not uni<strong>for</strong>m over all the surface. Leaching occurs at preferredcrystallographic sites, twin pl<strong>an</strong>es, micr<strong>of</strong>ractures, <strong>an</strong>d dislocations.An additional parameter not addressed in our study is the effect <strong>of</strong> permeabilityon <strong>albitization</strong>. Fine-grained s<strong>an</strong>dstone presumably has low initial permeability <strong>an</strong>dmuch <strong>of</strong> its original water could be lost through compaction/cementation by the time<strong>of</strong> <strong>albitization</strong> (Boles <strong>an</strong>d Coombs, 1977). Increasing <strong>albitization</strong> with increasing grainsize has been reported (Boles <strong>an</strong>d Coombs, 1977), but these results may indirectlyreflect permeability effects rather th<strong>an</strong> solely a surface area effect. Albitization requiresa local steady state composition <strong>of</strong> the pore fluid, in which Ca 2 <strong>an</strong>d Al 3 leave thelocal system on a const<strong>an</strong>t rate <strong>an</strong>d fluid tr<strong>an</strong>sport is not rapid. Differences inpermeability <strong>an</strong>d flow rates would establish different saturation states <strong>of</strong> the fluid overshort time intervals <strong>an</strong>d would ch<strong>an</strong>ge from one depth to <strong>an</strong>other dramatically. Ingeneral, s<strong>an</strong>dstone with contrasting permeability may have different <strong>an</strong>orthite contentthrough time <strong>an</strong>d different degrees <strong>of</strong> <strong>albitization</strong>.Effect <strong>of</strong> heating rates.—Heating rates c<strong>an</strong> influence the dissolution, precipitation, <strong>an</strong>deven the nucleation rates <strong>of</strong> mineral reactions (Luttge <strong>an</strong>d others, 1998). The effect <strong>of</strong> theheating rate on <strong>albitization</strong> c<strong>an</strong> be evaluated by comparing results from our simulations.For inst<strong>an</strong>ce the S<strong>an</strong> Joaquin Basin consists <strong>of</strong> relatively young deposits, from 6.7 to 33 my,but mostly less th<strong>an</strong> 20 my old, <strong>an</strong>d moderate to low thermal gradients between 25 <strong>an</strong>d27°C/km. At temperatures higher th<strong>an</strong> 170°C, our <strong>model</strong>ing suggest that at low temperaturealbite has replaced almost all plagioclase, leaving the <strong>an</strong>orthite mole fraction below 5percent. The Gulf Coast Tertiary Basin consists <strong>of</strong> sediments 30 to 35 my old <strong>an</strong>d arelatively high thermal gradient between 32 <strong>an</strong>d 37°C/km. Albitization occurs therebetween 90 <strong>an</strong>d 130°C, <strong>an</strong>d at temperatures higher th<strong>an</strong> 130°C the plagioclase approachespure albite composition. In contrast, the Denver Basin was subject to periods <strong>of</strong> 50 my attemperatures in the neighborhood <strong>of</strong> 90°C, <strong>an</strong>d, in our <strong>model</strong>, plagioclase has notreached pure albite composition as it does in the S<strong>an</strong> Joaquin <strong>an</strong>d Gulf Coast Basins. Basedon these three different sets <strong>of</strong> simulations, it is clear that <strong>albitization</strong> is primarily sensitiveto temperature rather th<strong>an</strong> time.The <strong>model</strong> dependence on temperature may also be evaluated by calculatingtime-temperature end-member values <strong>for</strong> <strong>albitization</strong>. In other words, calculatingminimum temperatures at const<strong>an</strong>t time (fig. 13A) <strong>an</strong>d minimum times at const<strong>an</strong>ttemperature (fig. 13B) required to albitize plagioclase. To per<strong>for</strong>m these calculations,we <strong>model</strong>ed <strong>albitization</strong> in fine-grain s<strong>an</strong>dstone, set the initial composition toAn 35 Ab 60 Or 5 , <strong>an</strong>d used our average <strong>kinetic</strong> parameters: E a <strong>an</strong>d A equal to 68 kJ/mole<strong>an</strong>d 6.5 10 3 1/cm 2 Ma.For the specified conditions, sedimentary rocks as old as 500 my must be subject totemperatures over 90°C in order to be albitized (fig. 10A). The minimum temperaturedecreases with time. S<strong>an</strong>dstone as old as 300 my must be subject to const<strong>an</strong>t temperaturesover 95°C, <strong>an</strong>d s<strong>an</strong>dstone 200 my old at temperatures over 105°C. In contrast, attemperatures over 170°C, <strong>albitization</strong> may occur in less th<strong>an</strong> 0.6 my (fig. 13B).It is import<strong>an</strong>t to emphasize that the replacement has a strong dependence ontemperature <strong>an</strong>d weak dependence on time (fig. 13). Nevertheless, these resultssuggest that <strong>albitization</strong> could occur slowly over long periods <strong>of</strong> time, <strong>an</strong>d might occurrapidly in rocks subject to rapid thermal events, <strong>for</strong> example, thermal alterationassociated with fluid flow along faults.some <strong>model</strong> implicationsWith available compositional <strong>an</strong>d petrologic data, our <strong>model</strong> may be used to putlimits on burial <strong>an</strong>d thermal histories <strong>of</strong> basins. The <strong>albitization</strong> <strong>kinetic</strong>s may be used asa paleo-thermometer providing independent calibration <strong>for</strong> other thermal indicators,such as vitrinite reflect<strong>an</strong>ce, TAI (Thermal Alteration Index, observed by sporedarkening), quartz precipitation overgrowth, <strong>an</strong>d smectite-illite, in sedimentary basins.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase339Fig. 13(A). Minimum temperatures required to complete <strong>albitization</strong> at const<strong>an</strong>t time assumingfine-grained s<strong>an</strong>dstone, which we considered to be <strong>an</strong> average value <strong>for</strong> all three basins. (B) Minimum timerequired to complete <strong>albitization</strong> at const<strong>an</strong>t temperature. We assume fine-grain size s<strong>an</strong>dstone, E a 68kJ/mole <strong>an</strong>d A 6.5 10 3 1/cm 2 Ma.The knowledge <strong>of</strong> the degree <strong>of</strong> <strong>albitization</strong>, <strong>for</strong> a given time-temperature data set,may be also useful <strong>for</strong> predicting the relative amounts <strong>of</strong> byproducts involved inreaction (2). For inst<strong>an</strong>ce, signific<strong>an</strong>t Ca 2 enrichment in pore fluids within thealbitized plagioclase zone has been reported (Fisher <strong>an</strong>d Boles, 1990). The enrichmentis mainly caused by the uptake <strong>of</strong> Na <strong>an</strong>d release <strong>of</strong> Ca 2 implied in reaction(2). The calcium release c<strong>an</strong> now be qu<strong>an</strong>titatively estimated in arkosic basins throughgeologic time with the <strong>albitization</strong> <strong>kinetic</strong> <strong>model</strong>. In short, the <strong>model</strong> has the potential


340 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>to lead to a better underst<strong>an</strong>ding <strong>of</strong> the evolution <strong>of</strong> basin fluids in geological timescales. The <strong>kinetic</strong> <strong>model</strong> may also be used by petroleum geologists <strong>for</strong> predicting theprecipitation <strong>of</strong> kaolinite or calcite (assuming a CO 2 supply) associated with <strong>albitization</strong>,during burial in a given oil reservoir. Any estimation may be done by simplyintegrating the <strong>kinetic</strong> equation through a time-temperature burial history <strong>of</strong> a givensiliciclastic <strong>for</strong>mation, <strong>an</strong>d applying basic rules <strong>of</strong> stoichiometry.conclusionsAlbitization <strong>of</strong> plagioclase trends present in sedimentary basins c<strong>an</strong> be reasonablyreproduced with a <strong>kinetic</strong> equation that has a rate const<strong>an</strong>t with a pre-exponentialparameter A (6.5 0.5) 10 3 1/cm 2 Ma, <strong>an</strong> activation energy E a 68 4 kJ/mole<strong>an</strong>d <strong>an</strong> weighted average saturation index log 10 2.632. The only fitted parameterwas A, the other parameters E a <strong>an</strong>d omega were fixed or calculated from real field dataprior to the <strong>kinetic</strong> <strong>an</strong>alysis. Our study demonstrates that time, surface area, initialcomposition, <strong>an</strong>d primarily temperature are the most import<strong>an</strong>t parameters controllingthe replacement reaction. Our <strong>model</strong> was tested against data from three geologicsettings: Texas Gulf Coast Basin, Denver Basin <strong>of</strong> Colorado, <strong>an</strong>d S<strong>an</strong> Joaquin Basin withdifferent burial <strong>an</strong>d thermal histories. These results evidence that the proposedequation <strong>an</strong>d its parameters c<strong>an</strong> reasonably reproduce the extent <strong>of</strong> the <strong>albitization</strong>within our knowledge <strong>of</strong> variables involved.The <strong>kinetic</strong> parameters yield rates within the r<strong>an</strong>ge <strong>of</strong> other diagenetic reactionsinvolving precipitation <strong>of</strong> silicates, specifically quartz overgrowth cementation, illiteprecipitation, <strong>an</strong>d with calculations <strong>of</strong> albite crystal growth rates. The initial composition<strong>of</strong> <strong>detrital</strong> plagioclase affects <strong>albitization</strong> during the initial stages. The grain sizeaffects <strong>albitization</strong> until the tr<strong>an</strong>s<strong>for</strong>mation is completed. Subtle differences in grainsize, initial composition, <strong>an</strong>d to a less extent the degree <strong>of</strong> order <strong>of</strong> the albitecomponent may be enough to explain the wide st<strong>an</strong>dard deviations reported in theliterature that we reference. We conclude that the <strong>albitization</strong> zone depends more ontemperature <strong>an</strong>d the heating rate, th<strong>an</strong> time <strong>an</strong>d fluid chemistry. All variables,however, are import<strong>an</strong>t <strong>an</strong>d must be considered.acknowledgmentsWe would like to acknowledge editorial work by Emma Perez <strong>an</strong>d Dr. ArthurSylvester, <strong>an</strong>d constructive comments by Drs. Thomas L. Dunn <strong>an</strong>d Bradley Hacker.Dr. Alicia Wilson provided import<strong>an</strong>t input with numerical methods <strong>an</strong>d some burialhistories from the S<strong>an</strong> Joaquin Basin. Comments <strong>an</strong>d revisions by Dr. Roll<strong>an</strong>dHellm<strong>an</strong>n <strong>an</strong>d the AJS <strong>an</strong>onymous reviewer greatly improved the content <strong>an</strong>d theoreticalbasis <strong>of</strong> our study. The U.S. Department <strong>of</strong> Energy (DOE) funded our researchunder gr<strong>an</strong>t no. 444033-22433.References Cited1984, Correlation <strong>of</strong> Stratigraphic Units <strong>of</strong> North America (COSUNA) Project Chart, Southern Cali<strong>for</strong>niaRegion: Tulsa, Oklahoma, The Americ<strong>an</strong> Association <strong>of</strong> Petroleum Geologists, chart 1.2002, Biostratigraphic chart <strong>for</strong> the Gulf <strong>of</strong> Mexico (OCS) Region, Department <strong>of</strong> Interior-MineralsM<strong>an</strong>agement Service, chart 1.Aagaard, P., <strong>an</strong>d Helgeson, H. C., 1982, Thermodynamic <strong>an</strong>d <strong>kinetic</strong> constraints on reaction rates amongminerals <strong>an</strong>d aqueous solutions, I. Theoretical considerations: Americ<strong>an</strong> Journal <strong>of</strong> Science, v. 282,p. 237–285.Aagaard, P., Egeberg, P. K., Saigal, G. C., Morad, S., <strong>an</strong>d Bjørlykke, K., 1990, Diagenetic <strong>albitization</strong> <strong>of</strong><strong>detrital</strong> K-feldspars in Jurassic, Lower Cretaceous <strong>an</strong>d Tertiary Clastic reservoir rocks from <strong>of</strong>fshoreNorway, II. Formation water chemistry <strong>an</strong>d <strong>kinetic</strong> considerations: Journal <strong>of</strong> Sedimentary Petrology,v. 60, p. 575–581.Anderson, G., <strong>an</strong>d Crerar, D., 1993, Thermodynamics in Geochemistry: The Equilibrium Model: New York,Ox<strong>for</strong>d University Press, 588 p.Awwiller, D. N., <strong>an</strong>d Summa, L. L., 1997, Quartz cement volume constrain on burial history <strong>an</strong>alysis: Anexample from the Eocene <strong>of</strong> Western Venezuela: Americ<strong>an</strong> Association <strong>of</strong> Petroleum Geologists <strong>an</strong>dSociety <strong>of</strong> Economic Paleontologists <strong>an</strong>d Mineralogists Annual Meeting, Abstracts, v. 6, p. 66.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase341Baccar, M. B., Fritz, B., <strong>an</strong>d Made, B., 1993, Diagenetic <strong>albitization</strong> <strong>of</strong> K-feldspars <strong>an</strong>d plagioclase ins<strong>an</strong>dstone reservoirs: thermodynamic <strong>an</strong>d <strong>kinetic</strong> <strong>model</strong>ing: Journal <strong>of</strong> Sedimentary Petrology, v. 63,p. 1100–1109.Bailey, E. B., <strong>an</strong>d Grabham, G. W., 1909, Albitization <strong>of</strong> basic plagioclase feldspars: Geological Magazine, v. 6,p. 250–256.Bethke, C. M., 1996, Geochemical Reaction Modeling: New York, Ox<strong>for</strong>d University Press, 397 p.Blum, A. E., <strong>an</strong>d Stillings, L. L., 1995, Feldspar dissolution <strong>kinetic</strong>s, in White, A. F., <strong>an</strong>d Br<strong>an</strong>tley, S. L.,editors, Chemical Weathering Rates <strong>of</strong> Silicate Minerals: Washington, D.C., Mineralogical Society <strong>of</strong>America Reviews in Mineralogy, v. 31, p. 291–351.Bodart, D. E., 1980, Genetic <strong>model</strong>s <strong>for</strong> the Precambri<strong>an</strong> Tarr albitite complex, southeastern Sinai:Geological Survey <strong>of</strong> Israel, p. 6–8.Boles, J. R., 1982, Active <strong>albitization</strong> <strong>of</strong> plagioclase, Gulf Coast Tertiary: Americ<strong>an</strong> Journal <strong>of</strong> Science, v. 282,p. 165–180.–––––– 1984, Secondary porosity reactions in the Stevens s<strong>an</strong>dstone, S<strong>an</strong> Joaquin Valley, Cali<strong>for</strong>nia, inMcDonald, D. A., <strong>an</strong>d Surdam, R. C., editors, Clastic Diagenesis: AAPG Memoir 37, p. 217–224.Boles, J. R., <strong>an</strong>d Coombs, D. S., 1977, Zeolite facies alteration <strong>of</strong> s<strong>an</strong>dstones in the Southl<strong>an</strong>d syncline, NewZeal<strong>an</strong>d: Americ<strong>an</strong> Journal <strong>of</strong> Science, v. 277, p. 982–1012.Boles, J. R., <strong>an</strong>d Fr<strong>an</strong>ks, S. G., 1979, Clay diagenesis in Wilcox s<strong>an</strong>dstones <strong>of</strong> southwest Texas: implications <strong>of</strong>smectite diagenesis on s<strong>an</strong>dstone cementation: Journal <strong>of</strong> Sedimentary Petrology, v. 49, p. 55–70.Boles, J. R., <strong>an</strong>d Ramseyer, K., 1988, Albitization <strong>of</strong> plagioclase <strong>an</strong>d vitrinite reflect<strong>an</strong>ce as paleothermalindicators, S<strong>an</strong> Joaquin Basin, in Graham, S. A., editor, Studies in the Geology <strong>of</strong> the S<strong>an</strong> Joaquin Basin:Pacific Section <strong>an</strong>d Society <strong>of</strong> Economic Paleontologists <strong>an</strong>d Mineralogists, v. 60, p. 129–139.Burnham, A. K., <strong>an</strong>d Sweeney, J. J., 1989, A chemical <strong>kinetic</strong> <strong>model</strong> <strong>of</strong> vitrinite maturation <strong>an</strong>d reflect<strong>an</strong>ce:Geochimica et Cosmochimica Acta, v. 53, p. 2649–2657.Burnham, A. K., Braun, R. L., Greg, H. R., <strong>an</strong>d Samoun, A. M., 1987, Comparison <strong>of</strong> methods <strong>for</strong> measuringkerogen pyrolisis rates <strong>an</strong>d fitting <strong>kinetic</strong> parameters: Energy <strong>an</strong>d Fuels, v. 1, p. 452–458.Cali<strong>for</strong>nia Oil <strong>an</strong>d Gas Fields, 1985, Central Cali<strong>for</strong>nia: Cali<strong>for</strong>nia Department <strong>of</strong> Conservation Division <strong>of</strong>Oil <strong>an</strong>d Gas, 3 rd edition, 1000 p.Callaway, D. C., 1971, Petroleum potential <strong>of</strong> the S<strong>an</strong> Joaquin Basin, Cali<strong>for</strong>nia, in Cram, I. H., editor, FuturePetroleum Provinces <strong>of</strong> the United States-Their geology <strong>an</strong>d potential: Americ<strong>an</strong> Association <strong>of</strong>Petroleum Geologists Memoir 15, p. 239–253.Comer, J., 1992, Thermal Alteration, in Pratt, L., Comer, J., <strong>an</strong>d Brassell, S., editors, Geochemistry <strong>of</strong> Org<strong>an</strong>icMatter in Sediments <strong>an</strong>d Sedimentary Rocks: SEPM Short Course 27: Tulsa, Oklahoma, Society <strong>for</strong>Sedimentary Geology, p. 73–100.Coombs, D. S., 1954, The nature <strong>an</strong>d alteration <strong>of</strong> some Triassic sediments from Southl<strong>an</strong>d, New Zeal<strong>an</strong>d:Royal Society New Zeal<strong>an</strong>d Tr<strong>an</strong>sactions, v. 82, p. 65–109.Davisson, M. L., <strong>an</strong>d Criss, R. E., 1996, Na-Ca-Cl relations in basinal fluids: Geochimica et CosmochimicaActa, v. 60, p. 2743–2752.Deer, W. A., Howie, R. A., <strong>an</strong>d Zussm<strong>an</strong>, J., 1963, Framework silicates: Rock-<strong>for</strong>ming minerals, v. 4: London,Longm<strong>an</strong>s, 435 p.Dewers, T., <strong>an</strong>d Ortoleva, P., 1990, Force <strong>of</strong> crystallization during the growth <strong>of</strong> siliceous concretions:Geology, v. 18, p. 204–207.Dodge, M. M., <strong>an</strong>d Posey, J. S., 1981, Structural cross sections, Tertiary <strong>for</strong>mations, Texas Gulf Coast: Austin,The University <strong>of</strong> Texas Austin, 6 p.Dumitru, T. A., 1988, Subnormal geothermal gradients in the Great Valley <strong>for</strong>earc basin, Cali<strong>for</strong>nia, duringFr<strong>an</strong>cisc<strong>an</strong> subduction; a fission track study: Tectonics, v. 7, p. 1201–1221.Elliot, C. W., <strong>an</strong>d Matis<strong>of</strong>f, G., 1996, Evaluation <strong>of</strong> <strong>kinetic</strong> <strong>model</strong>s <strong>for</strong> the smectite to illite tr<strong>an</strong>s<strong>for</strong>mation:Clays <strong>an</strong>d Clay Minerals, v. 44, p. 77–87.Eskola, P., Vuoristo, U., <strong>an</strong>d R<strong>an</strong>kama, K., 1935, An experimental illustration <strong>of</strong> spilite reaction: ComptesRendus de la Society geologique de Finl<strong>an</strong>de.Fisher, J. B., <strong>an</strong>d Boles, J. R., 1990, Water-rock interaction in Tertiary s<strong>an</strong>dstones, S<strong>an</strong> Joaquin Basin,Cali<strong>for</strong>nia, U.S.A.: Diagenetic controls on water composition: Chemical Geology, v. 83, p. 83–101.Fox, L. K., ms, 1989, Albitization <strong>of</strong> Jurassic plutons in the southern Bristol Mountains, east-central MojaveDesert, southeastern Cali<strong>for</strong>nia: Ph. D. thesis, University <strong>of</strong> Cali<strong>for</strong>nia, S<strong>an</strong>ta Barbara, 324 p.Gallagher, K., <strong>an</strong>d Ev<strong>an</strong>s, E., 1991, Estimating <strong>kinetic</strong> parameters <strong>for</strong> org<strong>an</strong>ic reactions from geological data:<strong>an</strong> example from the Gippsl<strong>an</strong>d basin, Australia: Applied Geochemistry, v. 6, p. 653–664.Galloway, W. E., Liu, K., Travis-Neuberger, D., <strong>an</strong>d Xue, L., 1994, Reference high-resolution correlation,cross sections, Paleogene section, Texas, Coastal plain: Austin, The University <strong>of</strong> Texas at Austin, 19 p.Garven, G., 1995, Continental-scale groundwater flow <strong>an</strong>d geological processes: Annual Review <strong>of</strong> Earth <strong>an</strong>dPl<strong>an</strong>etary Sciences, v. 23, p. 89–117.Gold, P. B., 1987, Textures <strong>an</strong>d geochemistry <strong>of</strong> authigenic albite from Miocene s<strong>an</strong>dstones, Louisi<strong>an</strong>a GulfCoast: Journal <strong>of</strong> Sedimentary Petrology, v. 57, p. 353–362.Hayes, M. J., <strong>an</strong>d Boles, J. R., 1992, Volumetric relations between dissolved plagioclase <strong>an</strong>d kaolinite ins<strong>an</strong>dstones: Implications <strong>for</strong> aluminum mass tr<strong>an</strong>sfer in the S<strong>an</strong> Joaquin Basin, Cali<strong>for</strong>nia, in Houseknecht,D. W., <strong>an</strong>d Pittm<strong>an</strong>, E. D., editors, Origin, Diagenesis, <strong>an</strong>d Petrophysics <strong>of</strong> Clay Minerals inS<strong>an</strong>dstones: Tulsa, Oklahoma, SEPM Special Publication, No. 47, p. 111–123.Helgeson, H. C., Murphy, W. M., <strong>an</strong>d Aagaard, P., 1984, Thermodynamic <strong>an</strong>d <strong>kinetic</strong> constraints on reactionrates among minerals <strong>an</strong>d aqueous solutions. II. Rate const<strong>an</strong>ts, effective surface area, <strong>an</strong>d thehydrolysis <strong>of</strong> feldspar: Geochimica et Cosmochimica Acta, v. 48, p. 2405–2432.Helgeson, H. C., Knox, A. m., Owens, C. E., <strong>an</strong>d Shock, E. L., 1993, Petroleum, oil field waters, <strong>an</strong>dauthigenic mineral assemblages: Are they in metastable equilibrium in hydrocarbon reservoirs?:Geochimica et Cosmochimica Acta, v. 57, p. 3295–3339.


342 R. J. Perez <strong>an</strong>d J. R. Boles—An <strong>empirically</strong> <strong>derived</strong>Hellm<strong>an</strong>n, R., 1994, The albite-water system: Part I. The <strong>kinetic</strong>s <strong>of</strong> dissolution as a function <strong>of</strong> pH at 100, 200<strong>an</strong>d 300 C: Geochimica et Cosmochimica Acta, v. 58, p. 595–611.Hess, H. H., 1950, Vertical mineral variation in the Great Dyke <strong>of</strong> Southern Rhodesia: Tr<strong>an</strong>sactionsGeological Society South Africa, v. 53, p. 159.Hu<strong>an</strong>g, W., Longo, J., <strong>an</strong>d Pevear, D., 1993, An experimentally <strong>derived</strong> <strong>kinetic</strong> <strong>model</strong> <strong>for</strong> smectite-to-illiteconversion <strong>an</strong>d its use as a geothermometer: Clays <strong>an</strong>d Clay Minerals, v. 41, p. 162–177.Johnson, J. W., Oelkers, E. H., <strong>an</strong>d Helgeson, H. C., 1992, SUPCRT92: A s<strong>of</strong>tware package <strong>for</strong> calculating thest<strong>an</strong>dard molal thermodynamic properties <strong>of</strong> minerals, gases, aqueous species, <strong>an</strong>d reactions from 1 to5000 bar <strong>an</strong>d 0 to 1000 C: Computers <strong>an</strong>d Geosciences, v. 18, p. 899–947.Kharaka, Y. K., <strong>an</strong>d Barnes, J. A., 1973, SOLMINEQ: solution-mineral equilibrium computations: U.S.Geological Survey Computer Contributions Report, p. 215–899.Knauss, K. G., <strong>an</strong>d Wolery, T. J., 1986, Dependence <strong>of</strong> albite solution <strong>kinetic</strong>s on pH <strong>an</strong>d time at 25 C <strong>an</strong>d 70C: Geochimica et Cosmochimica Acta, v. 50, p. 2481–2497.Labotka, T., Cole, D., Mostafa, F., <strong>an</strong>d Riciputi, L., 2002, Coupled cation <strong>an</strong>d oxygen exch<strong>an</strong>ge betweenalkali feldspar <strong>an</strong>d aqueous chloride solution: Goechimica et Cosmochimica Acta, v. 66, p. A427.Lachenbruch, A. H., <strong>an</strong>d Sass, J. H., 1980, Heat flow <strong>an</strong>d energetics <strong>of</strong> the S<strong>an</strong> Andreas fault zone: Journal <strong>of</strong>Geophysical Research, v. 85, p. 6185–6222.L<strong>an</strong>d, L. S., <strong>an</strong>d Milliken, K. L., 1981, Feldespar diagenesis in the Frio Formation, Brazoria County, TexasGulf Coast: Geology, v. 9, p. 314–318.L<strong>an</strong>der, R., <strong>an</strong>d Walderhaug, O., 1999, Predicting porosity through simulating s<strong>an</strong>dstone compaction <strong>an</strong>dquartz cement: Americ<strong>an</strong> Association <strong>of</strong> Petroleum Geologists Bulletin, v. 83, p. 433–449.Lasaga, A., 1984, Chemical <strong>kinetic</strong>s <strong>of</strong> water-rocks interactions: Journal <strong>of</strong> Geophysical Research, v. 89,p. 4009–4025.–––––– 1995, Fundamental approaches in describing mineral dissolution <strong>an</strong>d precipitation rates, in White,A. F., <strong>an</strong>d Br<strong>an</strong>tley, S. L., editors, Chemical Weathering Rates <strong>of</strong> Silicate Minerals: Washington, D.C.,Mineralogical Society <strong>of</strong> America, Reviews in Mineralogy <strong>an</strong>d Geochemistry, v. 31, p. 23–86.–––––– 1998, Kinetic theory in the earth sciences: Princeton, Princeton University Press, 811 p.Lopatin, N. V., 1971, Temperature <strong>an</strong>d geological time as factors <strong>of</strong> carbonifaction: Akademiya Nauk SSSRIzvestiya, Seriya Geologicheskaya, v. 3, p. 95–106.Luttge, A., <strong>an</strong>d Metz, P., 1991, Mech<strong>an</strong>ism <strong>an</strong>d <strong>kinetic</strong>s <strong>of</strong> the reaction: 1 dolomite 2 quartz 1 diopside 2CO 2 investigated by powder experiments: C<strong>an</strong>adi<strong>an</strong> Mineralogist, v. 4, p. 803–821.–––––– 1993, Mech<strong>an</strong>isms <strong>an</strong>d <strong>kinetic</strong>s <strong>of</strong> the reaction: 1 dolomite 2 quartz 1 diopside 2CO 2 .Acomparison <strong>of</strong> rock-sample <strong>an</strong>d powder experiments: Contributions to Mineralogy <strong>an</strong>d Petrology,v. 115, p. 155–164.Luttge, A., Neum<strong>an</strong>n, U., <strong>an</strong>d Lasaga, A., 1998, The influence <strong>of</strong> heating rate on the <strong>kinetic</strong>s <strong>of</strong>mineral reactions: An experimental study <strong>an</strong>d computer <strong>model</strong>s: Americ<strong>an</strong> Mineralogist, v. 83,p. 501–515.Mackenzie, A. S., <strong>an</strong>d Mackenzie, D. P., 1983, Isomerization <strong>an</strong>d aromatization <strong>of</strong> hydrocarbons insedimentary basins <strong>for</strong>med by extension: Geological Magazine, v. 220, p. 417–470.Maliva, R. G., <strong>an</strong>d Siever, R., 1988, Diagenetic replacement controlled by <strong>for</strong>ce <strong>of</strong> crystallization: Geology,v. 96, p. 688–691.Matthews, A., 1980, Influences <strong>of</strong> <strong>kinetic</strong>s <strong>an</strong>d mech<strong>an</strong>ism in metamorphism: a study <strong>of</strong> albite crystallization:Geochimica et Cosmochimica Acta, v. 44, p. 387–402.Merino, E., 1975a, Diagenesis in Tertiary s<strong>an</strong>dstones from Kettlem<strong>an</strong> North Dome, Cali<strong>for</strong>nia - II. Interstitialsolutions: distribution <strong>of</strong> aqueous species at 100 C <strong>an</strong>d chemical relation to the diagenetic mineralogy:Geochimica et Cosmochimica Acta, v. 39, p. 1629–1645.–––––– 1975b, Diagenesis in Tertiary s<strong>an</strong>dstones from Kettlem<strong>an</strong> North Dome, Cali<strong>for</strong>nia: I. Diageneticmineralogy: Journal <strong>of</strong> Sedimentary Petrology, v. 45, p. 320–336.Merino, E., <strong>an</strong>d Dewers, T., 1998 Implications <strong>of</strong> replacement <strong>for</strong> reaction-tr<strong>an</strong>sport <strong>model</strong>ing. Journal <strong>of</strong>Hydrology, v. 209, p. 137–146.Merino, E., Nahon, D., <strong>an</strong>d Yifeng, W., 1993, Kinetics <strong>an</strong>d mass tr<strong>an</strong>sfer <strong>of</strong> pseudomorphic replacement:application to replacement <strong>of</strong> parent minerals <strong>an</strong>d kaolinite by Al, Fe, <strong>an</strong>d Mn oxides duringweathering: Americ<strong>an</strong> Journal <strong>of</strong> Science, v. 293, p. 135–155.Middleton, G., 1972, Albite <strong>of</strong> secondary origin in Charny S<strong>an</strong>dstones, Quebec: Journal <strong>of</strong> SedimentaryPetrology, v. 42, p. 341–349.Moody, J. B., Jenkins, J. E., <strong>an</strong>d Meyer, D., 1985, An experimental investigation <strong>of</strong> the <strong>albitization</strong> <strong>of</strong>plagioclase: C<strong>an</strong>adi<strong>an</strong> Mineralogist, v. 23, p. 583–592.Morad, S., Morten, B., Knarud, R., <strong>an</strong>d Nystuen, J., 1990, Albitization <strong>of</strong> <strong>detrital</strong> plagioclase in Triassicreservoir s<strong>an</strong>dstones from the Snorre Field, Norwegi<strong>an</strong> North Sea: Journal <strong>of</strong> Sedimentary Petrology,v. 60, p. 411–425.Murray, G. E., 1960, Geological framework <strong>of</strong> Gulf Coastal Province <strong>of</strong> United States, in Shepard, F. P. a. o.,editor, Recent sediments, northwestern Gulf <strong>of</strong> Mexico: Tulsa, Oklahoma, Americ<strong>an</strong> Association <strong>of</strong>Petroleum Geologists Publication, p. 5–33.Nahon, D., <strong>an</strong>d Merino, E., 1997, Pseudomorphic replacement in tropical weathering: evidence,geochemical consequences, <strong>an</strong>d <strong>kinetic</strong>-rheological origin: Americ<strong>an</strong> Journal <strong>of</strong> Science, v. 297,p. 393–417.Oelkers, E. H., 1996, Physical <strong>an</strong>d Chemical Properties <strong>of</strong> Rock <strong>an</strong>d fluids <strong>for</strong> chemical mass tr<strong>an</strong>sportcalcultations, in Litchtner, P. C., Steefel, C. I., <strong>an</strong>d Oelkers, E. H., editors, Reactive Tr<strong>an</strong>sport in PorousMedia: Reviews in Mineralogy, Mineralogical Society <strong>of</strong> America, v. 34, p. 131–182.Oelkers, E. H., Bjorkum, P. A., <strong>an</strong>d Murphy, W. M., 1996, A petrographic <strong>an</strong>d computational investigation <strong>of</strong>quartz cementation porosity reduction in North Sea s<strong>an</strong>dstones: Americ<strong>an</strong> Journal <strong>of</strong> Science, v. 296,p. 420–452.


<strong>kinetic</strong> <strong>model</strong> <strong>for</strong> <strong>albitization</strong> <strong>of</strong> <strong>detrital</strong> plagioclase343Oelkers, E. H., Bjorkum, P. A., Walderhaug, O., Nadeau, P. H., <strong>an</strong>d Murphy, W. M., 2000, Making diagenesisobey thermodynamics <strong>an</strong>d <strong>kinetic</strong>s: the case <strong>of</strong> quartz cementation in s<strong>an</strong>dstones from <strong>of</strong>fshoremid-Norway: Applied Geochemistry, v. 15, p. 295–309.O’Neil, J. R., 1977, Stable isotopes in Mineralogy: Physics <strong>an</strong>d Chemistry <strong>of</strong> Minerals, v. 2, p. 105–123.O’Neil, J. R., <strong>an</strong>d Taylor, H. P., Jr., 1967, The oxygen isotope <strong>an</strong>d cation exch<strong>an</strong>ge chemistry <strong>of</strong> feldspars:Americ<strong>an</strong> Mineralogist, v. 52, p. 1414–1437.Perez, R. J., Chatellier, J. Y., <strong>an</strong>d L<strong>an</strong>der, R. H., 1999a, Use <strong>of</strong> quartz cementation <strong>kinetic</strong> <strong>model</strong>ing toconstrain burial histories. Examples from the Maracaibo basin, Venezuela: Revista Latinoameric<strong>an</strong>a deGeoquimica Org<strong>an</strong>ica, v. 5, p. 39–46.Perez, R. J., Gosh, S., Chatellier, J. Y., <strong>an</strong>d L<strong>an</strong>der, R. H., 1999b, Application <strong>of</strong> s<strong>an</strong>dstone diagenetic<strong>model</strong>ing to reservoir quality assessment <strong>of</strong> the Misoa Formation, Bachaquero Field, Maracaibo Basin,Venezuela: Annual Meeting Exp<strong>an</strong>ded Abstracts - Americ<strong>an</strong> Association <strong>of</strong> Petroleum Geologists,v. 1999, p. A107.Perkins, E. H., Kharaka, Y. K., Gunter, W. D., <strong>an</strong>d DeBraal, J. D., 1990, Geochemical <strong>model</strong>ing <strong>of</strong> water-rockinteractions using SOLMINEQ.88, in MELCHIOR, D. C., <strong>an</strong>d BASSETT, R. L., editors, ChemicalModeling <strong>of</strong> Aqueous Systems II: Americ<strong>an</strong> Chemical Society Symposium Series: Washington, D.C.,Americ<strong>an</strong> Chemical Society, p. 117–127.Pittm<strong>an</strong>, E. D., 1988, Diagenesis <strong>of</strong> Terry s<strong>an</strong>dstone (Upper Cretaceous), Spindle Field, Colorado: Journal <strong>of</strong>Sedimentary Petrology, v. 58, p. 785–800.–––––– 1989, Nature <strong>of</strong> the Terry s<strong>an</strong>dstone Reservoir, Spindle Field, Colorado, in Coalson, E. B., Editor,S<strong>an</strong>dstone Reservoirs: Denver, The Rocky Mountain Association <strong>of</strong> Geologists, p. 245–254.Putnis, A., 2002, Mineral replacement reactions: from macroscopic observations to microscopic mech<strong>an</strong>isms:Mineralogical Magazine, v. 66(5), p. 689–708.Pytte, A., <strong>an</strong>d Reynolds, R., 1988, The thermal tr<strong>an</strong>s<strong>for</strong>mation <strong>of</strong> smectite to illite, in Nasser, N., <strong>an</strong>dMcCulloh, T., editors, Thermal history <strong>of</strong> sedimentary basins: Berlin, Springer-Verlag, p. 133–140.Ramseyer, K., Boles, J. R., <strong>an</strong>d Lichtner, P. C., 1992, Mech<strong>an</strong>isms <strong>of</strong> Plagioclase Albitization: Journal <strong>of</strong>Sedimentary Petrology, v. 62, p. 349–356.Reyh<strong>an</strong>i, M. M., Freij, S., <strong>an</strong>d Parkinson, G. M., 1999, In situ atomic <strong>for</strong>ce microscope investigation <strong>of</strong> thegrowth <strong>of</strong> secondary nuclei produced by contact <strong>of</strong> different growth phases <strong>of</strong> potash alum crystalsunder supersaturated solutions: Journal <strong>of</strong> Crystal Growth, v. 198/99, p. 258–263.Robie, R. A., <strong>an</strong>d Waldbaum, D. R., 1968, Thermodynamic properties <strong>of</strong> minerals <strong>an</strong>d related subst<strong>an</strong>ces at298.15 K (25 C) <strong>an</strong>d one atmosphere (1.1013 bars) pressure <strong>an</strong>d at higher temperatures: GeologicalSurvey Bulletin, v. 1259, p. 1–256.Rosenbauer, R. J., Bish<strong>of</strong>f, J. L., <strong>an</strong>d Zierenberg, R. A., 1988, The laboratory <strong>albitization</strong> <strong>of</strong> mid-oce<strong>an</strong> ridgebasalt: Journal <strong>of</strong> Geology, v. 96, p. 237–244.Saxena, S. K., <strong>an</strong>d Ribbe, P. H., 1972, Activity-composition relations in feldspars: Contributions to Mineralogy<strong>an</strong>d Petrology, v. 37, p. 131–138.Siever, R., 1983, Burial history <strong>an</strong>d diagenetic reaction <strong>kinetic</strong>s: Americ<strong>an</strong> Association <strong>of</strong> PetroleumGeologist, Bulletin, v. 67, n. 4, p. 684–6911.Slaby, E., 1992, Ch<strong>an</strong>ges in the structural state <strong>of</strong> secondary albite during progressive <strong>albitization</strong>: NeuesJahbruch fuer Mineralogie, Monatshefte, v. 7, p. 321–335.Smith, J. V., 1972, Critical review <strong>of</strong> synthesis <strong>an</strong>d occurrence <strong>of</strong> plagioclase feldspars <strong>an</strong>d possible phasediagram: Journal <strong>of</strong> Geology, v. 80, p. 505–525.Stallard, M. L., <strong>an</strong>d Boles, J. R., 1989, Oxygen isotope measurements <strong>of</strong> albite-quartz-zeolite mineralassemblages, Hokonui Hills, Southl<strong>an</strong>d, New Zeal<strong>an</strong>d: Clay <strong>an</strong>d Clay Minerals, v. 37, p. 409–418.Tester, A. C., <strong>an</strong>d Atwater, G. I., 1934, The occurrence <strong>of</strong> authigenic feldspar in sediments: Journal <strong>of</strong>Sedimentary Petrology, v. 4, p. 23–31.Velde, B., <strong>an</strong>d Vasseur, G., 1992, Estimation <strong>of</strong> the diagenetic smectite to illite tr<strong>an</strong>s<strong>for</strong>mation in timetemperaturespace: Americ<strong>an</strong> Mineralogist, v. 77, p. 967–976.Walderhaug, O., 1994, Precipitation rates <strong>for</strong> quartz cement in s<strong>an</strong>dstones determined by fluid-inclusionmicrothermometry <strong>an</strong>d temperature-history <strong>model</strong>ing: Journal <strong>of</strong> Sedimentary Research, v. A64, p. 324–333.–––––– 1996, Kinetic <strong>model</strong>ing <strong>of</strong> quartz cementation <strong>an</strong>d porosity loss in deeply buried s<strong>an</strong>dstone reservoirs:Americ<strong>an</strong> Association <strong>of</strong> Petroleum Geologists Bulletin, v. 80, p. 731–745.Walther, J. V., <strong>an</strong>d Wood, B. J., 1984, Rate <strong>an</strong>d mech<strong>an</strong>ism in prograde metamorphism: Contributions toMineralogy <strong>an</strong>d Petrology, v. 88, p. 246–259.Waples, D. W., 1980, Time <strong>an</strong>d temperature in petroleum <strong>for</strong>mation: Application <strong>of</strong> Lopatin’s method topetroleum exploration: Americ<strong>an</strong> Association <strong>of</strong> Petroleum Geologist Bulletin, v. 64, p. 916–926.Wilson, A., Boles, J., <strong>an</strong>d Garven, G., 2000, Calcium mass tr<strong>an</strong>sport <strong>an</strong>d s<strong>an</strong>dstone diagenesis duringcompaction-driven flow: Stevens S<strong>an</strong>dstone, S<strong>an</strong> Joaquin Basin, Cali<strong>for</strong>nia: Geological Society <strong>of</strong>America Bulletin, v. 112, p. 845–856.Winkler, U., <strong>an</strong>d Luttge, A., 1999, The influence <strong>of</strong> CaCl 2 on the <strong>kinetic</strong>s <strong>of</strong> the reaction 1Tremolite 3Calcite 2Quartz 5Diopside 3CO 2 1H 2 O. An experimental investigation: Americ<strong>an</strong> Journal <strong>of</strong>Science, v. 299, p. 393–427.Wood, J. R., <strong>an</strong>d Boles, J. R., 1991, Evidence <strong>for</strong> episodic cementation <strong>an</strong>d diagenetic recording <strong>of</strong> seismicpumping events, North Coles Levee, Cali<strong>for</strong>nia, U.S.A.: Applied Geochemistry, v. 6, p. 509–521.Zhiyong, H., 2003, “Theoretical <strong>an</strong>d numerical basis <strong>for</strong> Genesis 1D basin <strong>model</strong>ing”: Downloaded J<strong>an</strong>uary28, http://www.zetaware.com/.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!