Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

begell house, inc.

JOURNAL PRODUCTION DEPARTMENT


50 North Street Date Proof Sent: November 7, 2017
Danbury, Connecticut 06810
203-456-6161 (Phone) Volume/Article ID#: Volume 20 ‐ JPM‐3398
203-456-6167 (Fax)
journals@begellhouse.com Total Pages: 11
Journal: Journal of Porous Media Year: 2017 Volume: 20

Article Title: Study of Dolomite Surface Stability by DFT Approach Considering Defects

Dear Raiza Hernández‐Bravo

Please review the attached PDF file which contains the author proof of your article.
This is your only opportunity to review the editing, typesetting, figure placement, and correctness of text,
tables, and figures. Answer copyeditor’s queries in the margin. Failure to answer queries will result in
the delay of publication of your article, so please make sure they are all adequately addressed. You
will not be charged for any corrections to editorial or typesetting errors; however, you will be
billed at the rate of $25 per hour of production time for rewriting, rewording, or otherwise revising
the article from the version accepted for publication (“author’s alterations”); any such charges will
be invoiced and must be paid before the article is published.
Please return your corrections clearly marked on the page proofs and email the corrected file or indicate your
corrections in a list, specifying the location of the respective revisions as precisely as possible. If you wish
to order offprints (see form below) please fill out and return the form with your corrections.
Please read the instructions carefully and email your corrections to journals@begellhouse.com, if possible
within 48 hours, not including weekends or holidays. If you need more time, please let us know at your earliest
convenience. No article will be published without confirmation of the author’s review. If we do not hear from
you within the allotted time, we will be happy to hold your article for a future issue, to give you more time to
make your corrections.
Attached is a form for ordering offprints, issues, or a subscription. If you wish to order extra issues or
offprints, please fill in the appropriate areas and fax the form or email it with your corrections. As
corresponding author, you will receive a complimentary PDF file, which is for your own personal use. This
file cannot be posted on any other website or used for distribution purposes.
Thank you for your assistance, and please reference J P M ‐ 3398 in your correspondence. Also, kindly
confirm receipt of your proofs.

Best Regards,

Contact for Author Proofs


Begell House Production
begell house, inc. Date: November 7, 2017
JOURNAL PRODUCTION DEPARTMENT Journal: Journal of Porous Media
50 North Street
Danbury, Connecticut 06810 Volume/Article ID: Volume 20 – JPM‐3398
203‐456‐6161 (Phone)
203‐456‐6167 (Fax) Article Title: Study of Dolomite Surface Stability by DFT Approach
journals@begellhouse.com Considering Defects
BILL TO: SHIP TO:





Dear Raiza Hernández‐Bravo
As corresponding author, you will receive a complimentary PDF file of your article. Please use the order form below to
order additional material and/or indicate your willingness to pay for color printing of figures (if applicable).
Begell House provides our author’s institution with a discount for subscriptions to the journal in which their article has been
published. For further information please contact Meghan Rohrmann at 1‐203‐456‐6161 or meghan@begellhouse.com
If placing an order, this form and your method of payment must be returned with your corrected page proofs. Please
include cost of shipment as indicated below; checks should be made payable to Begell House, Inc., and mailed to the above
address. If a purchase order is required, it may arrive separately to avoid delaying the return of the corrected proofs.

OFFPRINTS OF ARTICLE* WIRE TRANSFER


PAGE COUNT OF ARTICLE Bank: Citibank, N.A. Br # 619
(round off to highest multiple of 8)
Routing #: 22 11 726 10
QTY. 4 8 16 24 32
Account #: 12 55 463 407
25 72 115 151 187 223 Swift Code: CITI US 33
50 84 127 163 199 236
CREDIT CARD PAYMENT
100 108 193 254 314 375
200 156 327 435 544 652 CREDIT CARD #
300 205 459 616 773 930 NAME ON CREDIT CARD

*If your page count or quantity amount is not listed please AMEX/ VISA/MC/ DISC/ EURO/ EXP.
email a request for prices to journals@begellhouse.com
OTHER:
Black and White Offprints: Prices are quoted above
Offprint Color Pages: Add $3 per color page times the CORPORATE PURCHASE ORDER
quantity of offprints ordered P. O. # _
Shipping: Add 20% to black and white charge
PAYMENT BY CHECK
Offprint Qty: $ INCLUDE THE FOLLOWING INFO ON YOUR CHECK: Article Refer-
Color Pages for Offprints: $ ence # and Offprints/Color/Subscription
Shipping Charges: $ Make checks payable to Begell House, Inc.
COST FOR COLOR PAGES PRINTED IN JOURNAL AUTHOR DISCOUNTS
Figures provided in color will appear in color online at Authors are given a discount when ordering copies of the issue
no cost. in which their article appears.
Author’s Institution is given a discount on an Institutional Sub-
Price Per Color Page: $ 425.00 per printed page
scription to the journal in which the author’s article appears.
Number of Color Pages: CONTACT Meghan Rohrmann for details on these discounts,
Total Cost for Printed Color Pages: $ 203-456-6161 or meghan@begellhouse.com
REVISED 8-8-16
Journal of Porous Media, 20(11):1–11 (2017)

STUDY OF DOLOMITE SURFACE STABILITY BY DFT


APPROACH CONSIDERING DEFECTS

of
Raiza Hernández-Bravo
Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas, Mexico City, Mexico, E-mail:
raizahernandezbravo@gmail.com

Original Manuscript Submitted: 10/5/2016; Final Draft Received: 12/14/2016

ro
The stability of dolomite surfaces has been studied in the present work through a nonparameterized density functional
theory (DFT) with the aim to build a molecular model of a rock surface in contact with oil in a carbonate reservoir. The
rock surface was studied in its pristine form by considering defects such as cationic substitution and anionic vacancies.
Therefore a systematic molecular study was performed by the substitution of Ca+2 by Mg+2 and their influence in
the generation of vacancies at the CO2− 3 and cationic sites. The molecular simulations showed that Mg
+2
presented
the highest probability for spontaneous creation of vacancies; that is, the reactions where the substitution of Ca by Mg
rP
took place are favored. Therefore the stability is higher when Ca is substituted by Mg. Surface energy results showed
that the most stable pristine surface is the one corresponding to the [104] plane of the conventional dolomite cell. The
developed model can be used to favor the dolomitization to improve the reservoir quality by the generation of vacancies
and stability by the substitution of Ca by Mg incrementing the reservoir porosity in the order of 13%.

KEY WORDS: dolomite, DFT, surface energy, defects

1. INTRODUCTION
Dolomite has been used to improve the reservoir quality by increasing permeability in tight carbonate reservoirs,
o
as the dolomite is less reactive than calcite, and dolomites are also more resistant to porosity loss with depth tan
limestones (Broomhall and Allan, 1987; Dehghan et al., 2011). The dolomite is a mineral composed of calcium and
magnesium carbonate [CaMg (CO3 )2 ] produced when the substitution of calcium by magnesium in the limestone rock
CaCO3 by ion exchange takes place (Deer et al., 1966). It has a rhombohedral structure with three unit formulates and
space group R3, which is derived by reducing the symmetry of the calcite (R3c) for the arrangement of Ca and Mg in
th

alternating octahedral layers (Deer et al., 1966). There exist several basis studies on dolomite (Domingo et al., 2006;
de Leeuw and Parker, 1997, 2001), which characterize its properties through the determination of the surface energy
and quantum mechanics (Hossain et al., 2011). Yudin and Hughes (1994) in their study concluded that the inclusion
of a short-range Van der Waals interaction will lead to a system governed by an effective surface energy by studying
the surface energy of solids from a classical approach basing their study on the interactions of Van der Waals and
charge interactions. Titiloye et al. (1998) and Born and Huang (1954) found that the most stable surfaces for calcite
Au

and dolomite are the [104] and [100] planes, respectively. Titiloye et al. (1998) performed atomistic simulations of the
differences of their equilibrium morphology of the described composition. The atomistic simulation technique was
based on the Born model of solids (Born and Huang, 1954), which assumes that the ions in the crystal interact via long-
range electrostatic forces and short-range forces can be described using parameterized interatomic potential functions.
Wright et al. (2002) studied the structures plane [104] of calcite, dolomite, and magnesite surfaces under wet and dry
conditions by performing atomistic simulations. Then Born’s model of solids was used to define interatomic potential
functions to model the long- and short-range forces acting between atoms or ions in the solids. The authors found
that the most reactive anhydrous surfaces are calcite, and magnesite is the least reactive. These two articles used
the same calculation method, but not the same surface, causing low reliability of the published results by atomistic
simulation (Titiloye et al., 1998). Therefore, in the present work, atomistic simulations were performed employing

1091–028X/17/$35.00 © 2017 by Begell House, Inc. www.begellhouse.com 1


2 Hernández-Bravo

NOMENCLATURE

A surface area PWC plane wave calculation


Ca×Mg replacement of atom calcium for Surface-Ca calcium atom removed from the surface

of
one of magnesium surface Surface-CO3 carbonate atom removed from the surface
[h,k,l] crystallographic orientation; Surface-Mg magnesium atom removed from the surface
Miller indices [104], [100], and [110] U total energy
LNWC localized numeric wave calculation
Mg×Ca replacement of atom magnesium for Greek Symbol
one of calcium surface Γ surface energy

ro
quantum mechanics approaches to determine the most stable dolomite surface and to study the [104], [100], and [110]
surfaces, based on cationic and anionic substitutions on the dolomite surfaces (Guntram et al., 2001; Xu et al., 2010;
Drelich and Miller, 1995), with the aim to identify the changes that can affect the stability of dolomite structures. The
present work showed that by classical mechanics calculations, the most stable surface corresponds to [100] surface,
but employing the quantum mechanics approaches, it was found that the most stable surface corresponded to [104].
rP
Finally, it was observed that the described surfaces along with the [110] surface correspond to the most employed
surfaces in the classical mechanics calculations.

2. METHODOLOGY
2.1 Computation Details
The present calculations were carried out using the density functional theory (DFT) approach developed by Hohen-
berg and Kohn (1964) (see also Koch and Holthausen, 2001) employing the Cambridge Serial Total Energy Package
(CASTEP) code and Dmol3 of Material Studio software. In CASTEP, the optimization was performed with the
o
functional generalized gradient approximation (GGA) Perdew–Wang exchange and correlation functional (PW91)
(Perdew and Wang, 1992), then a converge threshold for the maximum energy change of 1.0 × 10−5 eV/atom and
for a maximum force of 0.03 eV/Å was fixed. The Vanderbilt ultrasoft pseudo potential (USP) method with a plane
wave basis and a cutoff of 340.0 eV was used for geometry optimization of the dolomite molecular structures and its
th

corresponding surfaces. The Monkhorst–Pack (MP) (Monkhorst and Pack, 1976; Pack and Monkhorst, 1977) mesh
of 3 × 3 × 2 was used to create the k-points within the irreducible segment of the Brillouin zone (BZ). Dmol3 soft-
ware code was employed to perform the geometry optimization with the functional GGA-PW91 (Perdew and Wang,
1992); therefore all electronic calculations were performed on the basis set of double numerical with D-functions
(DND) (Seminario et al., 1991). Also, a convergence threshold for the maximum energy change of 1.0 × 10−5 Ha
was used for a maximum force of 0.002 Ha.
In the work, also used was the PW91 functional because it meets the requirements for any functional correlation
Au

and exchange and is a nonparameterized functional, which does not depend on adjusted parameters of experimental
values and with no physical sense. Moreover, a pseudo potential was used to describe the inner electrons or core of
atoms which were obtained using the PW91; thus by using this for the valence electrons, inconsistencies are avoided.

2.2 Model Construction


The optimized dolomite structural parameters, used in the present calculations for the construction of the primitive
dolomite cell, have already been reported by Hossain et al. (2011) and will be used in the present work. Figure 1
shows the employed models for the bulk and [104], [100], [110] surfaces of dolomite calculations. For the dolomite
bulk mass optimization, there were employed the lattice parameters, which were previously reported by Hossain et

Journal of Porous Media


Study of Dolomite Surface Stability by DFT Approach 3

of
ro
FIG. 1: Molecular structure of bulk and surfaces dolomite studied

rP
al. (2011) (see Table 1), and it was done departing from the bulk of dolomite presented in the database of the software
Material Studio, then were performed the [100], [110], and [104] cuts, which correspond to Miller indices.

2.3 Surface Energy


For the determination of the surface energy, there exists two quantum methodologies named localized numeric plane
wave (LNWC) and plane wave calculation (PWC). In LNWC were used numerical functions, where the atomic basis
functions were obtained from the solution of DFT equations for individual atoms (Delley, 1988, 1990, 1996). In the
PWC were used pseudo potentials, which consist of inner electrons or the core of an atom, which are not significantly
affected by a change in the chemical environment, because of the strong attraction between these core electrons and
o
the atomic nucleus. In this case, only valence electrons were considered, which are usually responsible for chemical
processes (Clark et al., 2005; Payne et al., 1992).
The surface energy is given by
USlab − UBulk
γ= (1)
A
th

where USlab is the energy of the surface unit of the crystal, UBulk is the energy of an equal number of atoms of the
bulk crystal, and A is the surface area. It can be observed that Eq. (1) is widely used for the calculation of the surface
energy and was employed in the previous works performed by Wright et al. (2002) and de Leeuw and Parker (2001)
related to the study of solid surfaces performed by the simple model called slab material cutting certain thickness.
The slab model was constructed from the unit cell with an increment in the direction of the plane studied, creating
a vacuum molecule, which has been sufficiently large to avoid interactions between a slab and its replica in the
Au

TABLE 1: Lattice parameters used for each


dolomite surface studied
Models a (Å) b (Å)
Surface [104] 7.699 4.806
Surface [100] 4.877 16.285
Surface [110] 8.149 6.379
Note. Data adapted from Hossain et al. (2011).

Volume 20, Issue 11, 2017


4 Hernández-Bravo

direction of the vacuum. Normally, metals and covalent solids require fewer vacuum and ionic compounds, because
its interactions may be important at intermediate distances (Somorjai, 1994). Then a vacuum of 10 Å is a reasonable
value for any system, as was used in the present work.
The slab model involves having a number of sufficient atomic layers, as in the present case, and a correct descrip-
tion of the surface electronic properties (as the bulk). Then it was necessary to observe the thickness, measured in the

of
atomic layers, which is needed for convergence purposes, in a sensitive way, for the surface energy calculations. It is
important to mention that the surface energy, according to thermodynamic concepts, is a system that tends to reduce
its free energy on reaching the equilibrium state. In some cases, the stable state can be achieved by the reduction of
the surface energy system. The slab model was constructed as a surface model, where the slab is relaxed when the
surface chemistry is studied.
The studied indices [104], [100], and [110] of different dolomite surfaces are related as a plane within the Bravais
lattice, which is determined by the position of three noncollinear points. The usual method to appoint is by Miller

ro
indices. The indices were defined by the smallest vector reciprocal grid normal to the plane, thus a vector of the
reciprocal grid with [h,k,l] Miller indices (Ashcroft and Mermin, 1976).
The lattice parameters used for the determination of the area for the different surfaces are shown in the appendix.
The lowest obtained surface energy value corresponds to the more stable surface, as shown in the following
equation:
∆E = Ef − Ei (2)

vacancy.

3. RESULTS AND DISCUSSION


rP
where ∆E is energy difference, Ef is surface energy substituted, and Ei is surface energy with calcium or magnesium

3.1 Surface Energy of Dolomite


The results for dolomite and corresponding studied surfaces used in the LNWC and PWC are shown in Table 2. The
DFT calculations show that surface energies for the dolomite surfaces studied in the present work have the order
[110] > [100] > [104], where the [104] surface is the lowest obtained energy, corresponding to the most stable, and it
is consistent with that reported by Fenter et al. (2007) for dolomite, as experimentally confirmed by the well-known
o
fact that the crystal structure of the dolomite deposits predominantly in the [104] direction (Hoffman and Armbruster,
2010; Reeder and Wenk, 1983). The energy values obtained for the [104] surface using PWC are smaller than those
obtained via LNWC but do show the same stable surface preference and therefore suggest that the use of these
potentials to calculate surface energies is valid.
th

It has been reported that the stability of the surface of the dolomite is mainly due to the different types of surface
terminations: Mg and Ca-Mg (Titiloye et al., 1998). Figure 2 shows the stacking sequences for the [104] dolomite
surface, which dominates the predicted equilibrium morphology, as seen in Fig. 3, since it is located in the center
of the faces and [100] surface, and it is below the [104] surface. The surface [110] was not considered because it is
the least stable. The equilibrium morphology (Bravais, 1913; Friedel, 1907; Donnay and Harker, 1937) represents
the changes in surface stability and generates a list of faces (crystal planes), which can form a facet crystal. When
performing the study of equilibrium morphology, in some way, it is confirmed that the [104] surface is the most stable
Au

surface. Moreover, Miller planes in the [104] and [100] surfaces were created, as shown in Figs. 4 and 5. Figure 4
shows that the planes intersect the majority of the atoms of the dolomite surface in a more compact form. For the

TABLE 2: Dolomite surface energy


Models LNWC PWC
Surface [104] 1.072 0.900
Surface [100] 1.298 1.242
Surface [110] 1.471 1.596

Journal of Porous Media


Study of Dolomite Surface Stability by DFT Approach 5

of
FIG. 2: Stacking sequence surfaces of [104] dolomite

ro
o rP
FIG. 3: Predicted equilibrium morphology of dolomite showing [104] and [100] surfaces
th
Au

FIG. 4: Sets of parallel planes surfaces [104]

Volume 20, Issue 11, 2017


6 Hernández-Bravo

of
ro
rP
FIG. 5: Sets of parallel planes surfaces [100]

case of Fig. 5, the planes form a “sandwich” with the atoms of the structure of the dolomite, and only very few atoms
o
are oriented through the planes to the [100] surface. Then it can be observed that the [104] surface corresponds to a
packaging closer, producing a lower energy and therefore the highest stability.

3.2 Dolomite Nature


th

As described earlier, the structure of the dolomite is a rich structure of magnesium atoms, which is in contrast to the
calcite structure. It has been reported that the CO23− terminated faces dominate the stability of the dolomite surface
(Titiloye et al., 1998), but in the present work, it can be argued that this is not possible. This can be rationalized in
terms of the relative ionic radii of Ca+2 and Mg+2 , since the Mg+2 ion is smaller and thus has a shorter Mg-CO3 bond
distance than Ca+2 -CO3 . Therefore the first coordination shell of Mg is smaller than Ca. This leads to a significant
out-of-plane distortion of the carbonate groups (Wright et al., 2002). Figures 6–12 show the performed reactions
Au

and the energy difference substitutions, and the generated vacancies correspond to the negative values corresponding
to the most stable ones. Figures 6, 7, and 8 show the vacancies generated on the dolomite surface by a calcium,
magnesium, or carbonate atom. In this work, it was observed that CO23− groups can be detached more easily from the
dolomite structure (see Fig. 8), causing vacancies and an effect of instability into the system. The physical-chemical
behavior of a dolomite rock is presented in a matrix of an oil field, which has the peculiarity of being electrically
neutral due to the interchange of electronic charges on their surface expressed by Ca+2 · · · CO23− · · · Mg+2 · · ·
CO23− , facilitating the analysis of the system. To study the stability of the dolomite surface, the substitution of Ca+2
by Mg+2 , and vice versa, was analyzed with the aim to investigate the effect of replacing surface ions. Table 3 shows
the energies of reactants and products with different calculation accuracy levels (coarse, medium, and fine) in the
dolomite (present on the same dolomite) and typical configurations within a reservoir.

Journal of Porous Media


Study of Dolomite Surface Stability by DFT Approach 7

of
ro
FIG. 6: Vacancy generated on the dolomite surface by a cal- FIG. 7: Vacancy generated on the dolomite surface by a mag-
cium atom nesium atom
o rP
th

FIG. 8: Vacancy generated on the dolomite surface by a carbonate atom

The consequences of this substitution lead to a surface [104] with Ca+2 or Mg+2 ions and identify as affects in
nature, reactivity, and order of the crystallographic structure by dolomite. From the results shown in Table 3, it can be
seen that the ground state energy or more stable for the system is obtained when a calcium atom is replaced by one of
Au

magnesium, and from its counterpart, it is observed that the phenomenon that is less stable would be the system with
lower calcium.
For the surface energy calculation [see Eq. (2)] was found a range from −326.412 J/mol to −1735.369 J/mol for
substitutions of Mg+2 (as seen in Figs. 9 and 10), which are reactions that are thermodynamically favored. Moreover,
those results indicated that the reactions are favored when Mg+2 substitutes a Ca+2 molecule since the size of the
Mg+2 radius is smaller than the Ca+2 radius. Also, the electronegativity of Mg+2 is higher than Ca+2 , as clearly
shown in Tables 1 and 4. Then the Mg+2 ions prefer to remain in the surface, forming a dolomite layer. Table 4 shows
some physicochemical properties of calcium and magnesium (Chang, 2011; Castellan, 1983). The substitutions in
the surface by magnesium ions can cause formation of calcium carbonate crystal due to the junction of calcium
with carbonates which can be detached, and the energy value is shown in Fig. 8. Therefore the structure with CO3

Volume 20, Issue 11, 2017


8 Hernández-Bravo

TABLE 3: Energies of reactants and products (J/mol)


Calculation accuracy level Surface [104] Ca+2 Mg+2 CO23−
Energy of reactants
Coarse −44.088 × 106 −1773192.723 −521405.742 −687934.386
Medium −44.092 × 106 −1773192.438 −521405.338 −688186.769

of
Fine −44.094 × 107 −1773192.288 −521405.062 −688231.267
Calculation accuracy level Mg × Ca Ca × Mg Surface-Ca Surface-Mg Surface-CO3
Energy of products
Coarse −45.338 × 106 −42.835 × 106 −42.313 × 106 −43.564 × 106 −43.399 × 106
Medium −45.342 × 106 −42.839 × 106 −42.317 × 106 −43.569 × 106 −43.403 × 106

ro
Fine −45.342 × 106 −42.839 × 106 −42.319 × 106 −43.571 × 106 −43.405 × 106

o rP
FIG. 9: Mg × Ca substitution FIG. 10: Ca × Mg substitution

TABLE 4: Physicochemical properties of calcium and mag-


nesium
th

Properties Calcium Magnesium


Electronegativity 1.0 1.2
Atomic radius (Å) 1.97 1.6
Ionic radius (Å) 1.06 0.78
1.ªIonization energy (kJ/mol) 590 738
2.ªIonization energy (kJ/mol) 1145 1450
Au

Note. Data adapted from Chang (2011) and Castellan (1983).

terminated would not be stable, as mentioned earlier. These results are related with those performed in the study
of Purton et al. (2006), who investigated the order/disorder in dolomite and found that the order is very dependent
on the positions of the cations, particularly cations adopting homogeneous cation layers. This occurs because of the
difference of ionic radii of Ca (1.06 Å) and Mg (0.78 Å) so that any substitution of magnesium for calcium within a
layer dominated by the other can affect significantly the ideal positions of the cations in the dolomite structure.
Fisler et al. (2000) and Navrotsky et al. (1999) performed experimental studies of the difference in enthalpy and
calorimeters; in the case of the order/disorder in dolomite, it was found that in both cases, a difference of high energy

Journal of Porous Media


Study of Dolomite Surface Stability by DFT Approach 9

in a system was presented. The size of the energy gap between ordered and disordered phases is significant, and they
predicted that it can only be accessed at temperatures above 600 K. Therefore it can be observed that when Mg+2
substitutes to Ca+2 , it is favored by a negative value of the energy, but at the same time, this can produce a phase of
order/disorder in dolomite, which would be significant at temperatures above 600 K.
The energies for Ca+2 range from 271.993 J/mol to −1735.369 J/mol (see Figs. 10–12). In the case of Ca-rich

of
dolomites, these substitutions can be less reactive and less stable because of the presence of calcium excess, which
may be an important factor to determine the site distribution and concentration of impurities.
In addition, the substitution of Mg+2 by Ca+2 on the surface of the dolomite causes drastic changes in the
morphology of the structure, as suggested by Katz and Matthews (1977) and Reeder and Markgraf (1986). Recent
works studied the role of magnesium in the crystallization of monohydrocalcite (Rodrı́guez-Blanco et al., 2009, 2011,
2014). Therefore magnesium substitutions in the dolomite structure provided an increase in the Mg/Ca ratio, which
causes the crystallization, and thus it can be assumed that some dolomite or calcite deposits found in the geologic

ro
records that had been formed at high Mg/Ca ratios could be from a secondary origin and may have originally formed
through a metastable phase.
o rP
FIG. 11: Ca × Mg substitution
th
Au

FIG. 12: Mg × Ca substitution

Volume 20, Issue 11, 2017


10 Hernández-Bravo

4. CONCLUSIONS

In the present work, the crystal structure of dolomite was optimized through surface energy using the nonparam-
eterized DFT approach. Differences of electronegativity and ionic radii among magnesium and calcium favor the
substitution of Ca+2 by Mg+2 , causing changes in homogeneous layers of dolomite. Therefore the aim of the present
study was to obtain the vacancies or dolomite substitutions when into a reservoir are injected flow improvers and to

of
ensure that such injections do not disturb the systems in a way that modifies the dolomite structural configuration. In
addition, employing quantum calculations, the stability of the surface [104] was shown, which was originally by the
means of classical mechanics. The present modeling strategy was validated with data reported in the literature.

ACKNOWLEDGMENT

ro
The author is grateful for finances and permission to publish within project 143686 “Water Control in the Reservoir,”
Conacyt-Sener Hydrocarbons Program.

REFERENCES
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, XXXX: Brooks Cole, 1976.
rP
Born, M. and Huang, K., Dynamical Theory of Crystal Lattices, Oxford: Oxford University Press, 1954.
Bravais, A., Etudes Crystallographiques, Paris: Academie des Sciences, 1913.
Broomhall, R.W. and Allan, J.R., Regional Caprock-Destroying Dolomite on the Middle Jurassic to Early Cretaceous Arabian
Shelf, SPE Formation Evaluation, SPE-13697-PA, pp. 435–441, 1987.
Castellan, G.W., Physical Chemistry, 3rd ed., XXXX: Addison-Wesley, 1983.
Chang, R., Fundamentos de Quı́mica, New York: McGraw-Hill, 2011.
Clark, S.J., Segall, M.D., Pickard, C.J., Hasnip, P.J., Probert, M.I.J., Refson, K., and Payne, M.C., First principles methods using
CASTEP, Z. Kristallogr., vol. 220, pp. 567–570, 2005.
Deer, W.A., Howie, R.A., and Zussman, J., An Introduction to the Rock Forming Minerals, 1st ed., New York: John Wiley, 1966.
o
Dehghan, A.A., Kharrat, R., Ghazanfari, M.H., and Vossoughi, S., Quantifying the role of pore geometry and medium heterogene-
ity on heavy oil recovery during solvent/co-solvent flooding in water-wet systems, J. Porous Media, vol. 14, no. 4, pp. 363–373,
2011.
de Leeuw, N.H. and Parker, S.C., Atomistic simulation of the effect of molecular adsorption of water on the surface structure and
th

energies of calcite surfaces, J. Chem. Soc. Faraday Trans., vol. 93, no. 3, pp. 467–475, 1997.
de Leeuw, N.H. and Parker, S.C., Surface-water interactions in the dolomite problem, Phys. Chem. Chem. Phys., vol. 3, no. 15, pp.
3217–3221, 2001.
Delley, B., A scattering theoretic approach to scalar relativistic corrections on bonding, Int. J. Quantum Chem., vol. 69, no. 3, pp.
423–433, 1988.
Delley, B., An all-electron numerical method for solving the local density functional for polyatomic molecules, J. Chem. Phys.,
vol. 92, no. 1, pp. 508–517, 1990.
Au

Delley, B., Fast calculation of electrostatic in crystals and large molecules, J. Phys. Chem., vol. 100, no. 15, pp. 6107–6110, 1996.
Domingo, C., Loste, E., Gómez-Morales, J., Garcı́a-Carmona, J., and Fraile, J., Calcite precipitation by a high-pressure CO2
carbonation route, J. Supercritical Fluids, vol. 36, no. 3, pp. 202–215, 2006.
Donnay, J.H.D. and Harker, D., A new law of crystal morphology extending the law of bravais, Am. Min., vol. 22, no. 5, pp.
446–467, 1937.
Drelich, J. and Miller, J.D., A critical review of wetting and adhesion phenomena in the preparation of polymer-mineral composites,
Min. Metall. Process., vol. 12, no. 4, pp. 197–204, 1995.
Fenter, P., Zhang, Z., Park, C., Sturchio, N.C., Hu, X.M., and Higgins, S.R., Structure and reactivity of the Dolomite (104)–water
interface: New insights into the Dolomite problem, Geochim. Cosmochim. Acta, vol. 71, no. 3, pp. 566–579, 2007.

Journal of Porous Media


Study of Dolomite Surface Stability by DFT Approach 11

Fisler, D.K., Gale, J.D., and Cygan, R.T., A shell model for the simulation of rhombohedral carbonate minerals and their point
defects, Am. Min., vol. 85, no. 1, pp. 217–224, 2000.
Friedel, M.G., Etudes Sur la Loi de Bravais, Bull. Soc. Franc. Min., vol. 30, pp. 326–455, 1907.
Guntram, J., Higgins, S.R., Eggleston, C.M., Knauss, K.G., and Schmahl, W.W., Dissolution kinetics of magnesite in acidic
aqueous solution, a hydrothermal atomic force microscopy (HAFM) study: Step orientation and kink dynamics, Geochim.

of
Cosmochim. Acta, vol. 65, no. 23, pp. 4257–4266, 2001.
Hoffman, C. and Armbruster, T., Clinotobermorite, Ca5 [Si3 O8 (OH)]2 · 4 H2 O \p=n-\Ca5[Si6O17]•5H2O, a Natural C\p=n-
\S\p=n-\H(I) type cement mineral: Determination of the substructure, Z. Kristallogr. Crystalline Mater., vol. 212, no. 12, pp.
864–873, 2010.
Hohenberg, P. and Kohn, W., Inhomogeneous electron gas, Phys. Rev., vol. 136, no. 3B, pp. B864–B871, 1964.
Hossain, F.M., Dlugogorrski, B.Z., Kennedy, E.M., Belova, I.V., and Murch, G.E., First-principles study of the electronic, optical
and bonding properties in dolomite, Comput. Mater. Sci., vol. 50, no. 3, pp. 1037–1042, 2011.

ro
Katz, A. and Matthews, A., The dolomitization of CaCO3 : An experimental study at 252–295◦ C, Geochim. Cosmochim. Acta, vol.
41, no. 2, pp. 297–308, 1977.
Koch, W. and Holthausen, M.C., A Chemist’s Guide to Density Functional Theory, 2nd ed., Weinheim, Germany: Wiley, 2001.
Monkhorst, H.J. and Pack, J.D., Special point for brillouin-zone integrations, Phys. Rev. B, vol. 13, no. 12, pp. 5188–5192, 1976.
Navrotsky, A., Dooley, D., Reeder, R., and Brady, P., Calorimetric studies of the energetics of order-disorder in the system
Mg1−x Fex Ca(CO3 )2 , Am. Min., vol. 84, no. 10, pp. 1622–1626, 1999.
rP
Pack, J.D. and Monkhorst, H.J., Special point for brillouin-zone integrations—A reply, Phys. Rev. B, vol. 16, no. 4, pp. 1748–1749,
1977.
Payne, M.C., Teter, M.P., Allan, D.C., Arias, T.A., and Joannopoulos, J.D., Iterative minimization techniques for Ab Initio total-
energy calculations: Molecular dynamics and conjugate gradients, Rev. Mod. Phys., vol. 64, no. 4, pp. 1045–1097, 1992.
Perdew, J.P. and Wang, Y., Accurate and simple analytic representation of the electron-gas correlation energy, Phys. Rev. B, vol.
45, no. 23, pp. 13244–13249, 1992.
Purton, J.A., Allan, N.L., Lavrentiev, M., Todorov, I.T., and Freeman, C.L., Computer simulation of mineral solid solutions, Chem.
Geol., vol. 225, nos. 3-4, pp. 176–188, 2006.
Reeder, R. and Markgraf, S., High-temperature crystal chemistry of dolomite, Am. Min., vol. 71, pp. 795–804, 1986.
o
Reeder, R.J. and Wenk, H.R., Structure refinement of some thermally disordered Dolomites, Am. Min., vol. 68, pp. 769–776, 1983.
Rodrı́guez-Blanco, J.D., Shaw, S., and Benning, L.G., The real time kinetics and mechanisms of nucleation and growth of Dolomite
from solution, Geochim. Cosmochim. Acta, vol. 73, no. 13, p. A1111, 2009.
Rodrı́guez-Blanco, J.D., Bots, P., Roncal-Herrero, T., Shaw, S., and Benning L.G., The role of pH and Mg on the stability and
crystallization of Amorphous Calcium Carbonate, J. Alloy. Compd., vol. 536, pp. S477–S479, 2011.
th

Rodrı́guez-Blanco, J.D., Shaw, D., Bots, P., Roncal-Herrero, T., and Benning L.G., The role of Mg in the crystallization of mono-
hydrocalcite, Geochim. Cosmochim. Acta, vol. 127, pp. 204–220, 2014.
Seminario, J., Concha, M., and Politzer, P., Calculation of molecular and energies by a local density functional approach, Int. J.
Quantum Chem., vol. 40, no. S25, pp. 249–259, 1991.
Somorjai, G.A., Introduction to Surface Chemistry and Catalysis, New York: John Wiley, 1994.
Titiloye, J.O., de Leeuw, N.H., and Parker, S.C., Atomistic simulation of the differences between Calcite and Dolomite surfaces,
Au

Geochim. Cosmochim. Acta, vol. 62, no. 15, pp. 2637–2641, 1998.
Wright, K., Cygan, R.T., and Slater, B., Impurities and non-stoichiometry in the bulk and on the (1014) surface of Dolomite,
Geochim. Cosmochim. Acta, vol. 66, no. 14, pp. 2541–2546, 2002.
Xu, M., Hu, X., Knauss, K.G., Steven, R., and Higgins, R., Dissolution kinetic of Calcite from 50–70◦ C: An atomic force mi-
croscopy study under near-equilibrium conditions, Geochim. Cosmochim. Acta, vol. 74, no. 15, pp. 4285–4297, 2010.
Yudin, M. and Hughes, B.D., Surface energy of solids, Phys. Rev. B, vol. 49, pp. 5638–5642, 1994.

Volume 20, Issue 11, 2017

You might also like