center for nanoscale impact of mineral reactive surface ... storage...pyroxene group smectite...

1
Center for Nanoscale Control of Geologic CO 2 Abstract Nagaoka Reservoir Rock Mineral Surface Area Approximations BET-Based Grain-Size Specific Mineral Surface Area References We compiled a range of mineral surface area estimates to evaluate in our model, including total or specific surface area (SSA), and effective surface area (ESA) estimates. BET SA of pure minerals, literature values (thick black line) Image perimeter SSA (x) Image perimeter SSA with scaling factor (*) Adjusted GSA- smooth sphere ESA (o) Model results shown in blue (+) Assuming a spherical particle, the geometric surface area (GSA, in units of m 2 /g) can be calculated as a function of particle diameters according to the following equation: Stillings & Brantley 1995. Geochim. et Cosmochim. Act., 59, 1483-1496.; White & Peterson 1990. in Melchior et al. 1990, ACS Symposium, ACS. Washington, D.C.; Chiyonobu et al. 2013. Energy Proc., 37, 3546-3553.; Oelkers et al., 1994, Geochim. et Cosmochim. Act., 58, 2011-2024; Gudbrandsson et al., 2014. Geochim. et Cosmochim. Act., 139, 154-172.; Tester et al., 1994, Geochim. et Cosmochim. Act., 58, 2407-2420. Acknowledgements Evaluating Mineral Surface Area Approximations Using CrunchTope Impact of mineral reactive surface area approximations on predictions of mineral dissolution rates in a CO 2 injection experiment Elizabeth H. Mitnick 1 , Lauren E. Beckingham 2 , Shuo Zhang 1 , Carl I. Steefel 2 , Li Yang 2 , Marco Voltolini 2 , Alexander M. Swift 3 , Jonathan Ajo-Franklin 2 , David R. Cole 3 , Julie M. Sheets 3 , Saeko Mito 4 , Ziqiu Xue 4 , Donald J. DePaolo 2,1 1 University of California, Berkeley, 2 Lawrence Berkeley National Laboratory, 3 Ohio State University, 4 Research Institute of Innovative Technology for the Earth (RITE) Injection of supercritical CO 2 into porous reservoirs, or geologic carbon sequestration, is a promising means of reducing atmospheric CO 2 emissions. However, the rate and extent of reactions between injected fluid and surrounding geology is not well understood. Continuum-scale modeling has emerged as a tool to predict in situ mineral reaction rates, although a major challenge associated with this approach lies in the uncertainty associated with mineral reactive surface areas. The present study is aimed at evaluating the impact of mineral reactive surface area approximations on predictions of reaction rates in a powder dissolution experiment. Reservoir samples from the Nagaoka pilot CO 2 injection site in Japan were reacted with CO 2 -acidified brine (pCO 2 = 100 bars) in a flowthrough reactor at 50° C and the effluent chemistry from the reactor was measured. The multicomponent reactive transport code CrunchTope is used to model both steady state and time-dependent effluent chemistry as a means of evaluating classical and novel image-based approximations of surface area. Reactive surface area approximations evaluated include physical surface areas, i.e., geometric and specific, and reactive-site weighted surface areas. In the simulations, multiple pools of the major reactive phases with different grain sizes and thus specific surface area were considered. Results indicate that the use of BET-based grain-size specific mineral surface areas are the most accurate way of matching observed mineral dissolution rates in the well-stirred powder experiments, although these conclusions may not apply equally to core samples where the pore structure is intact. This work was supported as part of the Center for Nanoscale Control of Geologic CO(NCGC) , an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science. GSA i = SurfaceArea i total mass i = 4π r 2 4 3 π r 3 ρ = 3 rρ d = particle diameter and ρ = mineral density The geometrical surface area is related to the BET surface area (BET) through a surface roughness factor (SRF): GSA i = 6 d ρ BET = GSA SRF Time (hours) 0 100 200 300 400 500 600 Concentration (mol/kg) x 10 -4 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 Ca BET low BET high IPSSA, Smectite BET IPSSA, Smectite Mito et al. SF ESA Exp Data Model Time (hours) 0 100 200 300 400 500 600 Concentration (mol/kg) x 10 -5 0 1 2 3 4 5 6 7 8 Mg BET low BET high IPSSA, Smectite BET IPSSA, Smectite Mito et al. SF ESA Exp Data Model Time (hours) 0 100 200 300 400 500 600 Concentration (mol/kg) x 10 -4 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 Si BET low BET high IPSSA, Smectite BET IPSSA, Smectite Mito et al. SF ESA Exp Data Model Mineral Grain Sizes (m) 10 -8 10 -7 10 -6 10 -5 10 -4 10 -3 model literature Quartz Feldspar Group Pyroxene Group Smectite Biotite Chlorite Serpentine Calcite Amphibole Surface Area Estimate (m 2 /g) 10 -3 10 -2 10 -1 10 0 10 1 10 2 10 3 Quartz Feldspar Group Pyroxene Group Smectite Biotite Amphibole Chlorite Serpentine Calcite *Chiyonobu et al. 2013 Rock samples are from well cores from the CO 2 injection test site in Nagaoka, Japan. Sample depth = 1093 m Mineral compositions in the sample determined by QEMSCAN analysis are entered into the reactive transport model’s thermodynamic database. Literature values of mineral dissolution rate constants were compiled exclusively from studies using flowthrough reactors under controlled T,P conditions. How does Crunch handle surface area inputs? Percentage of total feldspar- grouped phases (%) Phase description Ideal Composition Composition using QEMSCAN 6.6 Anorthite CaAl2Si2O8 - 59.4 Labradorite (Ca,Na)(Al(Al,Si)Si2O8) (Ca0.62Na0.38)(Al(Al0.62 Si0.38)Si2O8) 28.5 Albite NaAlSi3O8 - 5.5 Sanidine KAlSi3O8 - Percentage of total pyroxene- grouped phases (%) Phase description Ideal Composition Composition using QEMSCAN 75.7 Ferrosilite FeSiO3 - 20.4 Augite (Ca,Na)(Mg,Fe,Al)(Si,Al)2O6 (Ca0.98Na0.02)(Mg0.65Fe0.25Al0.1)(Si0.96Al0.04)2O6 3.5 Epidote Ca2(Al2Fe 3+ )(Si2O7) (SiO4)(O)(OH) - 0.3 Others - - r = kA 1 Q K " # $ % & ' M ( ) * * + , - - n Advanced surface area estimates do not consistently over- or under-predict effluent concentrations of major cations. Only simulations with image-based and high literature BET SA approximations approach time-dependent dissolution behavior. Best model fits to powder dissolution experiment data are attained with BET- based grain size specific formulations of surface area. To determine if these estimates are the best approximations of surface area overall, they need to be tested on material with intact pore structure. 1. SSA (m 2 /g) x molar weight of mineral (g/mol) 2. Divide by molar volume (mol/m 3 ) m 2 mineral/m 3 mineral. 3. Multiply by volume fraction (m 3 mineral/m 3 porous media), m 2 mineral/m 3 media. 4. This value A gets used in rate equation (TST). Summary of findings Feldspar and Pyroxene Group Mineral Compositions from QEMSCAN Analyses Literature BET grain size ranges (squares) compared with estimated grain sizes based on image analysis and model-fit SA (circles)

Upload: others

Post on 25-Aug-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Center for Nanoscale Impact of mineral reactive surface ... storage...Pyroxene Group Smectite Biotite Amphibole Chlorite Serpentine Calcite *Chiyonobu et al. 2013 Rock samples are

C e n t e r f o r N a n o s c a l e Control of Geologic CO2

Abstract

Nagaoka Reservoir Rock

Mineral Surface Area Approximations

BET-Based Grain-Size Specific Mineral Surface Area

References

We compiled a range of mineral surface area estimates to evaluate in our model, including total or specific surface area (SSA), and effective surface area (ESA) estimates. BET SA of pure minerals, literature values (thick black line)

Image perimeter SSA (x)

Image perimeter SSA with scaling factor (*)

Adjusted GSA- smooth sphere ESA (o)

Model results shown in blue (+)

Assuming a spherical particle, the geometric surface area (GSA, in units of m2/g) can be calculated as a function of particle diameters according to the following equation:

Stillings & Brantley 1995. Geochim. et Cosmochim. Act., 59, 1483-1496.; White & Peterson 1990. in Melchior et al. 1990, ACS Symposium, ACS. Washington, D.C.; Chiyonobu et al. 2013. Energy Proc., 37, 3546-3553.; Oelkers et al., 1994, Geochim. et Cosmochim. Act., 58, 2011-2024; Gudbrandsson et al., 2014. Geochim. et Cosmochim. Act., 139, 154-172.; Tester et al., 1994, Geochim. et Cosmochim. Act., 58, 2407-2420.

Acknowledgements

Evaluating Mineral Surface Area Approximations Using CrunchTope

Impact of mineral reactive surface area approximations on predictions of mineral dissolution rates in a CO2 injection experiment

Elizabeth H. Mitnick1, Lauren E. Beckingham2, Shuo Zhang1, Carl I. Steefel2, Li Yang2, Marco Voltolini2, Alexander M. Swift3, Jonathan Ajo-Franklin2, David R. Cole3, Julie M. Sheets3, Saeko Mito4, Ziqiu Xue4, Donald J. DePaolo2,1

1University of California, Berkeley, 2Lawrence Berkeley National Laboratory, 3Ohio State University, 4Research Institute of Innovative Technology for the Earth (RITE)

Injection of supercritical CO2 into porous reservoirs, or geologic carbon sequestration, is a promising means of reducing atmospheric CO2 emissions. However, the rate and extent of reactions between injected fluid and surrounding geology is not well understood. Continuum-scale modeling has emerged as a tool to predict in situ mineral reaction rates, although a major challenge associated with this approach lies in the uncertainty associated with mineral reactive surface areas. The present study is aimed at evaluating the impact of mineral reactive surface area approximations on predictions of reaction rates in a powder dissolution experiment. Reservoir samples from the Nagaoka pilot CO2 injection site in Japan were reacted with CO2-acidified brine (pCO2 = 100 bars) in a flowthrough reactor at 50° C and the effluent chemistry from the reactor was measured. The multicomponent reactive transport code CrunchTope is used to model both steady state and time-dependent effluent chemistry as a means of evaluating classical and novel image-based approximations of surface area. Reactive surface area approximations evaluated include physical surface areas, i.e., geometric and specific, and reactive-site weighted surface areas. In the simulations, multiple pools of the major reactive phases with different grain sizes and thus specific surface area were considered. Results indicate that the use of BET-based grain-size specific mineral surface areas are the most accurate way of matching observed mineral dissolution rates in the well-stirred powder experiments, although these conclusions may not apply equally to core samples where the pore structure is intact.

This work was supported as part of the Center for Nanoscale Control of Geologic CO₂ (NCGC) , an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science.

GSAi =SurfaceAreai

total

massi=4πr243 πr

3ρ=3rρ

d = particle diameter and ρ = mineral density The geometrical surface area is related to the BET surface area (BET) through a surface roughness factor (SRF):

GSAi =6dρ

BET =GSA∗SRF

Time (hours)0 100 200 300 400 500 600

Con

cent

ratio

n (m

ol/k

g)

x 10-4

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2Ca

BET lowBET highIPSSA, Smectite BETIPSSA, Smectite Mito et al.SFESAExp DataModel

Time (hours)0 100 200 300 400 500 600

Conc

entra

tion

(mol

/kg)

x 10-5

0

1

2

3

4

5

6

7

8Mg

BET lowBET highIPSSA, Smectite BETIPSSA, Smectite Mito et al.SFESAExp DataModel

Time (hours)0 100 200 300 400 500 600

Conc

entra

tion

(mol

/kg)

x 10-4

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8Si

BET lowBET highIPSSA, Smectite BETIPSSA, Smectite Mito et al.SFESAExp DataModel

Mineral Grain Sizes (m)10-8 10-7 10

-610-5 10-4 10-3

model

literature

Quartz

Feldspar Group

Pyroxene Group

Smectite

Biotite

Chlorite

Serpentine

Calcite

Amphibole

Surface Area Estimate (m2 /g)10-3 10-2 10-1 100 101 102 103

Quartz

Feldspar Group

Pyroxene Group

Smectite

Biotite

Amphibole

Chlorite

Serpentine

Calcite

*Chiyonobu et al. 2013

Rock samples are from well cores from the CO2 injection test site in Nagaoka, Japan. Sample depth = 1093 m

Mineral compositions in the sample determined by QEMSCAN analysis are entered into the reactive transport model’s thermodynamic database. Literature values of mineral dissolution rate constants were compiled exclusively from studies using flowthrough reactors under controlled T,P conditions.

How does Crunch handle surface area inputs?

Percentage of total

feldspar-grouped

phases (%)

Phase description Ideal Composition Composition using QEMSCAN

6.6 Anorthite CaAl2Si2O8 - 59.4 Labradorite (Ca,Na)(Al(Al,Si)Si2O8) (Ca0.62Na0.38)(Al(Al0.62 Si0.38)Si2O8) 28.5 Albite NaAlSi3O8 - 5.5 Sanidine KAlSi3O8 -

Percentage of total

pyroxene-grouped

phases (%)

Phase description Ideal Composition Composition using QEMSCAN

75.7 Ferrosilite FeSiO3 -

20.4 Augite (Ca,Na)(Mg,Fe,Al)(Si,Al)2O6 (Ca0.98Na0.02)(Mg0.65Fe0.25Al0.1)(Si0.96Al0.04)2O6

3.5 Epidote Ca2(Al2Fe3+)(Si2O7) (SiO4)(O)(OH) - 0.3 Others - -

r = kA 1− QK"

#$

%

&'M(

)**

+

,--

n

Advanced surface area estimates do not consistently over- or under-predict effluent concentrations of major cations. Only simulations with image-based and high literature BET SA approximations approach time-dependent dissolution behavior. Best model fits to powder dissolution experiment data are attained with BET-based grain size specific formulations of surface area. To determine if these estimates are the best approximations of surface area overall, they need to be tested on material with intact pore structure.

1.  SSA (m2/g) x molar weight of mineral (g/mol) 2.  Divide by molar volume (mol/m3) à m2 mineral/m3 mineral. 3.  Multiply by volume fraction (m3 mineral/m3 porous media),

à m2 mineral/m3 media. 4.  This value A gets used in rate equation (TST).

Summary of findings

Feldspar and Pyroxene Group Mineral Compositions from QEMSCAN Analyses

Literature BET grain size ranges (squares) compared with estimated grain sizes based on image analysis and model-fit SA (circles)