You are on page 1of 30

Subscriber access provided by AUSTRALIAN NATIONAL UNIV

Article
Room temperature magnesite precipitation
Ian M. Power, Paul Alexander Kenward, Gregory Martin Dipple, and Mati Raudsepp
Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.7b00311 • Publication Date (Web): 06 Oct 2017
Downloaded from http://pubs.acs.org on October 6, 2017

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Crystal Growth & Design is published by the American Chemical Society. 1155
Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 29 Crystal Growth & Design

1
2
3
4
5
Room temperature magnesite precipitation
6
7 Ian M. Power*,1, Paul A. Kenward2, Gregory M. Dipple2, and Mati Raudsepp2
8 1
School of the Environment, Trent University, Peterborough, ON K9L 0G2, Canada. IMP –
9
10
ianpower@trentu.ca
2
11 Department of Earth, Ocean and Atmospheric Sciences, The University of British Columbia,
12 2207 Main Mall, Vancouver, British Columbia V6T 1Z4, Canada; PAK – thepkone@gmail.com,
13 GMD – gdipple@eos.ubc.ca, MR – mraudsepp@eos.ubc.ca
14
15
Abstract. Magnesite (MgCO3) is one of the most stable sinks for carbon dioxide (CO2)
16
17 and is therefore of great interest for long-term carbon storage. Although magnesite is the
18 thermodynamically stable form of magnesium carbonate, the kinetic inhibition of low-
19 temperature precipitation has hindered the development of carbon sequestration strategies that
20 can be economically conducted under ambient temperature. Here we document the precipitation
21 of magnesite from waters (magnesite saturation index = 1.45) in batch reactors at room
22
23
temperature with the aid of carboxylated polystyrene microspheres over the course of 70 days.
24 Microspheres provide surfaces with a high density of carboxyl groups that act to bind and
25 dehydrate Mg2+ ions in solution, thereby minimizing the kinetic barrier and facilitating
26 magnesite formation. Magnesite crystals are observed on sphere surfaces and their organic
27 matrices. Mineral identification was confirmed by X-ray diffraction and selected area electron
28
diffraction of a thin section obtained by focused ion beam milling. We demonstrate that kinetic
29
30 barriers to magnesite formation can be overcome at ambient conditions. Incorporating surfaces
31 with high carboxyl site densities into ex situ mineral carbonation processes and the use of such
32 ligands for deep geologic CO2 storage may offer novel and economically viable strategies for
33 permanent carbon storage.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 *corresponding author: Ian M. Power, ianpower@trentu.ca
58
59
60 1
ACS Paragon Plus Environment
Crystal Growth & Design Page 2 of 29

1
2
3
4
5
Room temperature magnesite precipitation
6
7 Ian M. Power*,1, Paul A. Kenward2, Gregory M. Dipple2, and Mati Raudsepp2
8 1
School of the Environment, Trent University, Peterborough, ON K9L 0G2, Canada. IMP –
9
10
ianpower@trentu.ca
2
11 Department of Earth, Ocean and Atmospheric Sciences, The University of British Columbia,
12 2207 Main Mall, Vancouver, British Columbia V6T 1Z4, Canada; PAK – thepkone@gmail.com,
13 GMD – gdipple@eos.ubc.ca, MR – mraudsepp@eos.ubc.ca
14
15
*corresponding author: Ian M. Power, ianpower@trentu.ca
16
17
18
19 ABSTRACT
20
21 Magnesite (MgCO3) is one of the most stable sinks for carbon dioxide (CO2) and is
22
23
24 therefore of great interest for long-term carbon storage. Although magnesite is the
25
26 thermodynamically stable form of magnesium carbonate, the kinetic inhibition of low-
27
28
temperature precipitation has hindered the development of carbon sequestration strategies that
29
30
31 can be economically conducted under ambient temperature. Here we document the precipitation
32
33 of magnesite from waters (magnesite saturation index = 1.45) in batch reactors at room
34
35
36
temperature with the aid of carboxylated polystyrene microspheres over the course of 70 days.
37
38 Microspheres provide surfaces with a high density of carboxyl groups that act to bind and
39
40 dehydrate Mg2+ ions in solution, thereby minimizing the kinetic barrier and facilitating
41
42
43 magnesite formation. Magnesite crystals are observed on sphere surfaces and their organic
44
45 matrices. Mineral identification was confirmed by selected area electron diffraction of a thin
46
47 section obtained by focused ion beam milling. We demonstrate that kinetic barriers to magnesite
48
49
50 formation can be overcome at ambient conditions. Incorporating surfaces with high carboxyl site
51
52 densities into ex situ mineral carbonation processes and the use of such ligands for deep geologic
53
54
CO2 storage may offer novel and economically viable strategies for permanent carbon storage.
55
56
57
58
59
60
2
ACS Paragon Plus Environment
Page 3 of 29 Crystal Growth & Design

1
2
3
1. INTRODUCTION
4
5
6 Mechanisms of carbonate formation have long been studied to further our understanding of
7
8 carbonate depositional environments, yet this area of research has gained greater attention
9
10
11
because of its relevance to permanent storage of carbon dioxide (CO2) in both ex situ mineral
12
13 carbonation processes and deep geologic storage.1-3 Ex situ mineral carbonation technologies
14
15 aim to capitalize on the globally abundant natural deposits of ultramafic and mafic rocks by
16
17
18 forming magnesite (MgCO3) as a stable sink for CO2.1 These proposed technologies are usually
19
20 operated at high temperature and pressure to achieve fast carbonation rates; however, these
21
22 approaches are not yet economically viable with current technology.3 At low temperatures,
23
24
25 carbonation of Mg-bearing feedstock, either abiotically or biologically, results in the
26
27 precipitation of metastable hydrated Mg-carbonate minerals such as nesquehonite
28
29
(MgCO3·3H2O), dypingite [Mg5(CO3)4(OH)2·5H2O] and hydromagnesite
30
31
32 [Mg5(CO3)4(OH)2·4H2O].4-9 These metastable phases require considerably more onerous
33
34 chemical conditions (e.g., higher Mg concentration) than those required to reach magnesite
35
36
37 saturation.10 This conundrum in balancing carbonation rates with acceptable costs has spurred a
38
39 wealth of scientific inquiry with the goal of developing economically feasible carbonation
40
41 pathways.
42
43
44 Magnesite and dolomite [CaMg(CO3)2] have precipitation rates that are orders of
45
46 magnitude slower than calcite (CaCO3) at Earth’s surface conditions, which has hampered the
47
48 formation of these minerals in the laboratory at room temperature.9, 11-13 These sluggish rates
49
50
51 have hindered our understanding of magnesite formation in natural environments and have
52
53 generally restricted carbon sequestration technologies that form magnesite to high temperature
54
55
56
systems.3, 11 However, magnesite has successfully been precipitated at 35 °C using supercritical
57
58
59
60
3
ACS Paragon Plus Environment
Crystal Growth & Design Page 4 of 29

1
2
3
CO2.14-17 At ambient pressure and low temperatures (<60 °C), magnesite precipitation is
4
5
6 kinetically inhibited due to the strong hydration of magnesium (Mg2+) ions in solution.10, 18 This
7
8 is the same root cause behind the dolomite problem, which is defined by the disparity between
9
10
11
the abundance of dolomite in the geologic record and its paucity in modern environments despite
12
13 being supersaturated in seawater. The “problem” is one of kinetics, and given sufficient time,
14
15 precipitation rates are adequate to explain the occurrence of dolomite as well as magnesite in the
16
17
18 geologic record without mediation or acceleration.19
19
20 Several studies have succeeded in increasing rates of disordered or ordered dolomite
21
22 precipitation at low temperature to observe its formation at the laboratory timescale and to begin
23
24
25 to demystify the processes that may accelerate its formation in sedimentary environments.20-23
26
27 The dehydration of Mg2+ ions is considered the rate-limiting step for the precipitation of both
28
29
dolomite and magnesite at low temperature.11 Kenward et al.21 demonstrated that carboxyl
30
31
32 functional groups on microbial cell walls could bind and partially dehydrate Mg2+ ions in an
33
34 energetically favorable process that facilitates dolomite precipitation. This mechanism was
35
36
37 validated with the use of densely carboxylated microspheres to abiotically precipitate dolomite
38
39 from solution at 30 °C.22 In the present study we show the formation of magnesite at room
40
41 temperature with the aid of carboxylated polystyrene microspheres in solutions slightly
42
43
44 supersaturated with respect to magnesite. Manipulation of this reaction pathway to accelerate
45
46 magnesite formation at low temperature may provide an economical alternative for carbon
47
48 storage, as there likely would be substantially less energy required in comparison to high
49
50
51 temperature and pressure systems.
52
53
54
55
56
2. EXPERIMENTAL SECTION
57
58
59
60
4
ACS Paragon Plus Environment
Page 5 of 29 Crystal Growth & Design

1
2
3
2.1. Batch experiments. Batch reactors were used to examine the precipitation of magnesite
4
5
6 at room temperature (~22 °C). Experimental solutions (100 mL of deionized water) were
7
8 composed of 10 mM NaHCO3 and 10 mM MgCl2·6H2O with an initial starting pH measured at
9
10
11
8.2. The solution chemistry had a magnesite saturation index (SI) of 1.45 as calculated using
12
13 PHREEQC V.324 with the LLNL database. Solutions were undersaturated with respect to
14
15 hydromagnesite (SI = -1.91) and nesquehonite (SI = -1.57). Solutions were filtered (0.22 µm)
16
17
18 and dispensed into serum vials that were sealed using butyl rubber stoppers and Al-crimp seals to
19
20 prevent degassing and evaporation. Low (0.7 COO-/Å2; 0.82 µm diameter) and high (10 COO-
21
22 /Å2; 20.3 µm diameter) carboxyl site density polystyrene microspheres were purchased from
23
24
25 Bangs Laboratories Inc. These spheres were composed of 99.90% polystyrene and ≤0.10%
26
27 sodium dodecyl sulfate. Reactors were amended with approximately 105 spheres/mL, and
28
29
experiments were run in triplicate over five time intervals for each experimental condition.
30
31
32 Reactors were not continuously mixed, and therefore most microspheres settled to the bottom.
33
34 Analyses of solutions prior to and after microsphere addition showed that bulk water chemistry
35
36
37 (pH and [Mg]) was the same within measurement error. Reactor vessels were opened and
38
39 sampled in succession, starting at 24 days of reaction with additional reactors being sampled
40
41 every 12 days for a total of 72 days. Control systems contained either no microspheres or
42
43
44 contained microspheres in deionized water. Millipore Swinnex filter holders (25 mm) and
45
46 IsoporeTM membrane filters (0.2 or 0.4 µm) were used to collect spheres and any precipitates.
47
48 While remaining in filter holders, solids were rinsed using filtered deionized water to prevent
49
50
51 formation of evaporite minerals. These solids were then dried at room temperature prior to
52
53 imaging and analyses, which were done within 2 days.
54
55
56
57
58
59
60
5
ACS Paragon Plus Environment
Crystal Growth & Design Page 6 of 29

1
2
3
2.2. Aqueous chemistry. A portable Thermo Scientific® Orion 4-Star pH/ISE probe was
4
5
6 calibrated and used to measure pH of experimental solutions before and after microsphere
7
8 addition. The total aqueous Mg concentration of solutions was determined using inductively
9
10
11
coupled plasma optical emission spectrometry (ICP-OES) employing a Varian 725-ES Optical
12
13 Emission Spectrometer, with a detection limit of 0.2 mg/L. Reproducibility was better than 5
14
15 mg/L based on repeated analysis of standards and duplicate samples. Cation samples were
16
17
18 acidified to 2% ultrapure HNO3 immediately after sampling.
19
20
21
22 2.3. Scanning electron microscopy. Samples were imaged at the Centre for High-
23
24
25 Throughput Phenogenomics at The University of British Columbia and Nanofabrication Facility
26
27 at The University of Western Ontario. Spheres and precipitates were immobilized onto EMD
28
29
Millipore IsoporeTM membrane filters (0.2 or 0.4 µm) and flushed with deionized-distilled water
30
31
32 to prevent evaporite minerals from forming. Filters were air dried before mounting onto
33
34 aluminum stubs using 12 mm carbon adhesive tabs. Samples were coated with either 8 nm of
35
36
37 iridium using a Leica EM MED020 coating system or 3 nm layer of osmium metal using a
38
39 Filgen osmium plasma coater (OPC 80T). A LEO (Zeiss) 1540 XB field emission scanning
40
41 electron microscope (SEM) was used to produce high-resolution images at an operating voltage
42
43
44 of 1.0 kV. Operating the SEM at a voltage of 10 kV, an Oxford Instruments’ INCAx-sight
45
46 energy dispersive spectrophotometer (EDS) was utilized for elemental analysis. Focused ion
47
48 beam (FIB) milling was used to obtain a thin section (~100 nm thickness) of the mineral
49
50
51 precipitate. At the University of British Columbia, samples were imaged using an FEI Helios
52
53 NanoLab 650 operating at 1.0 kV voltage.
54
55
56
57
58
59
60
6
ACS Paragon Plus Environment
Page 7 of 29 Crystal Growth & Design

1
2
3
2.3. X-ray diffraction. Mineral phases from experiments were identified using X-ray
4
5
6 diffraction (XRD) data. Samples were mounted as slurries on to zero diffraction quartz plates
7
8 with anhydrous ethanol and allowed to dry at room temperature. All XRD data were collected
9
10
11
using a Bruker D8 Focus Bragg-Brentano diffractometer with CoKα radiation and a step size of
12
13 0.02° over a range of 30-70°2θ at 2.0 s/step. Fe monochromator foil, 0.6 mm divergence slit,
14
15 incident and diffracted beam Soller slits, and a Lynx Eye detector were used. A long fine focus
16
17
18 Co X-ray tube was operated at 35 kV and 40 mA using a take-off angle of 6°. Search-match
19
20 software, DIFFRACplus EVA 14,25 and the International Centre for Diffraction Database PDF-
21
22 4+ 2010 were used for phase identification.
23
24
25
26
27 2.4. Transmission electron microscopy. Imaging of the thin section was performed at The
28
29
University of British Columbia's Bioimaging Facility. The FEI Tecnai G2 transmission electron
30
31
32 microscope (TEM) was operated at 200 kV with a LaB6 filament capable of 1.4 Å resolution.
33
34 The Tecnai TEM is equipped with a high-tilt, high precision motorized stage and a high-speed
35
36
37 AMT 2K side mount CCD camera, and a high-resolution FEI Eagle 4K bottom mount CCD
38
39 camera for capturing digital images. Crystal orientation and reflections shown in the selected
40
41 area electron diffraction (SAED) pattern were identified using CrystalMaker® v2.7 and
42
43
44 SingleCrystal v2.2. An electron diffraction pattern of an aluminum standard from Ted Pella Inc.
45
46 (Prod. No. 619) was used to calculate lattice spacings for an unknown mineral precipitate within
47
48 a FIB thin section (Table S1).
49
50
51
52
53 3. RESULTS AND DISCUSSION
54
55
56
57
58
59
60
7
ACS Paragon Plus Environment
Crystal Growth & Design Page 8 of 29

1
2
3
Experiments using low carboxyl site density microspheres yielded few mineral
4
5
6 precipitates, none of which distinctly resembled magnesite (Fig. 1A). The precipitate in Figure
7
8 1A did contain Mg, C, and O based on EDS analysis. Likewise, no mineral precipitates were
9
10
11
identified in the control experiments that lacked spheres or which contained spheres in deionized
12
13 water (see Supplementary Material Fig. S1A). Conversely, after 60 days of reaction, numerous
14
15 (<3 µm) mineral precipitates were observed in two of the three replicate experiments that used
16
17
18 the high carboxyl site density spheres. Precipitates appeared to be well crystalline and had
19
20 rhombohedral crystal morphologies. Crystal sizes typically ranged from 200 nm to 2 µm and
21
22 were in close association with microspheres and their organic matrices (Fig. 1B-D; see Fig. S1
23
24
25 for additional micrographs). Precipitates had portions that were rhombohedral in shape, yet also
26
27 had rough step-like features, indicative of spiral crystal growth.11 Figure 1C shows numerous
28
29
grains of various sizes held within an organic matrix (presumably made of styrene) hanging from
30
31
32 a sphere. Mineral precipitates near, but not in direct contact with spheres, may have been
33
34 detached during sampling (Fig. 1D). Precipitates on filter paper (Fig. 1D) were almost always
35
36
37 associated with organic matrices similar to those attached to spheres (Fig. 1C). Importantly, the
38
39 mineral precipitates had markedly similar appearance to magnesite grains that have formed in the
40
41 temperature range of 1-11 °C within modern hydromagnesite-magnesite playas (Fig. 1E).26, 27
42
43
44 Very similar magnesite crystal morphologies have been noted by others in high temperature
45
46 experiments.18, 28, 29 Hydrated Mg-carbonate minerals such as lansfordite (MgCO3·5H2O; short,
47
48 prismatic), nesquehonite (elongated, prismatic), dypingite [Mg5(CO3)4(OH)2·5H2O; rosettes,
49
50
51 flaky], hydromagnesite (rosettes, platy) and artinite [Mg2CO3(OH)2·3H2O; acicular] have very
52
53 different crystals morphologies than the rhombohedral crystals of magnesite.26, 30-34 Thus, the
54
55
56
57
58
59
60
8
ACS Paragon Plus Environment
Page 9 of 29 Crystal Growth & Design

1
2
3
crystal morphologies of the experimental precipitates (Fig. 1) are strongly indicative of
4
5
6 magnesite.
7
8 Elemental analyses using EDS revealed that the precipitated crystals contained Mg, C, and
9
10
11
O. This semi-quantitative technique measured an average Mg:C molar ratio of 0.8:1 (n = 22),
12
13 which is similar to magnesite (Fig. 1F-G). Of the hydrated Mg-carbonate minerals, only
14
15 lansfordite and nesquehonite have a 1:1 Mg:C molar ratio, whereas dypingite and
16
17
18 hydromagnesite have a ratio of 5:4 and artinite has a much greater ratio of 2:1. The elemental
19
20 composition and Mg:C molar ratio further support that the precipitated crystals are magnesite.
21
22 Based on SEM observations, the polystyrene spheres appeared to compose nearly 100% of
23
24
25 the total solids collected. Detection limits for magnesium carbonate minerals by XRD are
26
27 typically greater than ~0.5 wt.%.35 Thus, conventional XRD analysis was hampered by the
28
29
minuscule abundance of mineral precipitates and obstruction by the X-ray amorphous
30
31
32 microspheres. Only samples from experiments with the most abundant mineral precipitates
33
34 produced XRD diffraction patterns with identifiable peaks (Fig. S2). XRD patterns of solids
35
36
37 from this experiment after 60 days only showed the most intense diffraction peak for magnesite
38
39 (1 0 4; d = 2.746 Å). The second (45% intensity) and third (35% intensity) most intense peaks at
40
41 50.3° and 63.5° 2θ may be hidden behind the signal from the amorphous polystyrene spheres.
42
43
44 A thin section was obtained using FIB milling (Fig. 2A-B) and imaged using TEM (Fig.
45
46 2C) with SAED patterns being collected (Fig. 2D and Fig. S3) to obtain a better mineral
47
48 identification. Lattice spacings were calculated using an electron diffraction pattern of an
49
50
51 aluminum standard (Fig. S3A) collected under the same conditions as the pattern of the unknown
52
53 mineral. The formula, K = Sd, was used to calculate a constant (K) that incorporates beam
54
55
56
wavelength, camera length, and associated variable crystallographic data, S values are ring
57
58
59
60
9
ACS Paragon Plus Environment
Crystal Growth & Design Page 10 of 29

1
2
3
diameters (cm) of the diffraction pattern of the aluminum standard, and d is the lattice spacing in
4
5
6 Å. An average K value for the first five rings was used to calculate the lattice spacings of the
7
8 unknown precipitate. In the [3 0 1] direction, there were reflections from the 1 2 3, 1 1 3, 2 1 6,
9
10
11
and 0 3 0 planes with measured lattice spacings of 2.10, 2.11, 1.70, and 2.34 Å, respectively.
12
13 Importantly, the measurement of these lattice spacings was done independently from their
14
15 identification using the software. These lattice spacings very closely match the known lattice
16
17
18 spacing of the aforementioned planes for magnesite (Table 1). The electron diffraction data,
19
20 crystal morphology, and chemical composition confirmed the precipitate to be magnesite.
21
22
23
24
25 3.1. Precipitation mechanisms. To our knowledge, we are the first to demonstrate the
26
27 formation of magnesite, either by direction precipitation from solution or transformation from a
28
29
precursor, at room temperature. This outcome was achieved by utilizing surfaces dense in
30
31
32 carboxyl groups to overcome the kinetic barrier associated with hydration of Mg2+ ions and thus
33
34 accelerating magnesite precipitation. No other precipitates were detected or observed prior to
35
36
37 identification of magnesite after 60 days of reaction. At this point, there were no signs of
38
39 intergrowth between magnesite and a hydrated Mg-carbonate mineral, such as hydromagnesite,
40
41 that would suggest formation from a precursor.36 As such, direct precipitation from solution is
42
43
44 the likely pathway for magnesite formation in this study.
45
46 Carbonate precipitation rates are related to the exchange rate of water molecules in the
47
48 metal coordination hydration shell, which is five orders of magnitude faster for Ca2+ ions than
49
50
51 the smaller, electron dense Mg2+ ions.12, 37 Magnesium ions have a high charge density and
52
53 stable hydration spheres; the inner sphere being comprised of six water molecules in octahedral
54
55
56
coordination.38, 39 The outer sphere is formed from 12 water molecules hydrogen-bonded to the
57
58
59
60
10
ACS Paragon Plus Environment
Page 11 of 29 Crystal Growth & Design

1
2
3
inner sphere.40 The net charge of a hydrated magnesium ion is ~1.2+ indicating significant
4
5
6 charge transfer to the surrounding water molecules.41 The energy cost for removal of a water
7
8 molecule from the outer and inner hydration spheres has been measured to be 5-12 and 20-80
9
10
11
kcal/mol, respectively.38 The sluggish exchange of water molecules of [Mg(H2O)6]2+ results in
12
13 precipitation rates of magnesite and dolomite being six and four orders of magnitude slower than
14
15 that of calcite at 25 °C, respectively.11, 12
16
17
18 The inner sphere binding of [Mg(H2O)6]2+ to a ligand may occur when the free energy of
19
20 the substitution reaction is negative.42 For example, the exchange of water for formate ions
21
22 (HCOO-), the simplest of carboxyl group compounds, is thermodynamically favorable for the
23
24
25 first three substitutions, yet unfavorable for a fourth substitution.42 Thus, ligands act to disrupt
26
27 the inner hydration shell and remove water molecules from Mg2+. Ligands can also affect the
28
29
composition of carbonate minerals by preferentially binding a competing cation. For instance,
30
31
32 carboxylated organic acids can increase the magnesium content of amorphous calcium carbonate,
33
34 a calcite precursor, due to their stronger affinity for Ca2+ in comparison to Mg2+.43
35
36
37 Microbial cell walls are host to a variety of functional groups, including carboxyl and
38
39 hydroxyl groups, giving these surfaces a net negative charge.44 Partly for this reason, microbes
40
41 are thought to play an important role in precipitating dolomite and its precursors in the natural
42
43
44 environment.20, 21 and could similarly influence magnesite precipitation. Regarding low-
45
46 temperature dolomite precipitation, Roberts et al.22 emphasize the importance of surfaces with
47
48 high carboxyl site densities. Nonmetabolizing archaea, Haloferax sulfurifontis, with carboxyl
49
50
51 site density of 0.1 COO-/Å2 were able to induce precipitation of dolomite.21 As an example of
52
53 organomineralization, Zhang et al.23 synthesized disordered dolomite at room temperature from
54
55
56
solutions containing either carboxymethyl cellulose or agar. Microbes and organic matter are
57
58
59
60
11
ACS Paragon Plus Environment
Crystal Growth & Design Page 12 of 29

1
2
3
ubiquitous to carbonate-forming environments and may provide carboxylated surfaces for
4
5
6 dehydrating Mg2+ ions in nature.21, 45 For instance, there is evidence of a microbial role in
7
8 magnesite formation in ancient lacustrine hypersaline environments;46 however, clearly the
9
10
11
conditions for low-temperature magnesite precipitation are rare given the scarcity of modern
12
13 deposits.26
14
15 The polystyrene microspheres used by Roberts et al.22 and in this study have a carboxyl
16
17
18 site density that is far greater than that of an average bacterium (0.06 COO-/Å2) found in nature
19
20 and H. sulfurifontis that was used to induce dolomite precipitation.22 Greater carboxyl site
21
22 density may partly explain the success of the high carboxyl site density spheres (10 COO-/Å2)
23
24
25 and failure of the low carboxyl site density spheres (0.7 COO-/Å2) for precipitating magnesite.
26
27 In addition to site density, the competitiveness of a ligand is also likely to affect carbonate
28
29
precipitation rates and mineralogy of precipitates. Ligand competitiveness is partly determined
30
31
32 by its charge, electron density distribution, polarizability, size, and rigidity relative to water that
33
34 is binding to the cation.42 Carboxylates have been shown to affect the relative abundances of
35
36
37 nesquehonite and magnesite formed through the reaction of forsterite with supercritical CO2
38
39 (scCO2) at 50 °C and 90 bar.47 Divalent carboxylates, including malonate [CH2(COO)22-] and
40
41 oxalate [(COO)22-], can form bidentate complexes that may better disrupt the inner hydration
42
43
44 shell relative to monovalent anions, such as acetate (CH3COO-), that forms monodentate
45
46 complexes. Trivalent citrate [C3H5O(COO)33-] was most effective at dehydrating Mg2+ ions,
47
48 possibly due in part to the affinity of its hydroxyl group for cations.47 Polycarboxylates may be
49
50
51 more effective in dehydrating Mg2+ ions in a similar fashion as having a surface with a high
52
53 carboxyl site density.
54
55
56
Molecular dynamics simulations investigating the effect of Mg-H2O interaction on
57
58
59
60
12
ACS Paragon Plus Environment
Page 13 of 29 Crystal Growth & Design

1
2
3
dolomite crystal growth revealed that Mg is thermodynamically easily incorporated into a calcite
4
5
6 crystal, but then inhibits incorporation of Ca unless water is removed from the surface.48
7
8 Consequently, a reduction in water activity may aid in incorporating magnesium ions. For
9
10
11
instance, Xu et al.49 were able to form amorphous magnesium carbonate and increase magnesium
12
13 substitution into calcite precipitated in formamide, a nonaqueous solvent. In our experiments,
14
15 magnesite crystals were commonly observed with an organic matrix (e.g., Fig. 1C); the volume
16
17
18 of which was likely reduced under ultrahigh vacuum in the SEM. This organic matrix may have
19
20 created localized areas of reduced water activity due to organic-water interactions that assisted
21
22 crystal growth.
23
24
25
26
27 3.2. Implications for CO2 sequestration. Modern (i.e., Holocene epoch) sedimentary
28
29
magnesite-forming environments are rare, only forming under unique conditions such as those
30
31
32 found in freshwater Mg-carbonate playas that are associated with ultramafic rocks26, 30 and
33
34 highly evaporative coastal environments such as the Coorong saltwater lagoons in South
35
36
37 Australia.50, 51 The scarcity of magnesite-forming systems is in stark contrast to the vast Ca-
38
39 carbonate deposits on Earth, which contain roughly five million times more carbon in
40
41 comparison to annual global greenhouse gas emissions;1, 3 despite magnesite (SI = 0.73) being
42
43
44 more supersaturated in seawater than calcite (SI = 0.53). Presumably, the Earth’s carbon cycle
45
46 would be dramatically different if magnesite formed with the same ease as calcium carbonate—
47
48 deposits of magnesite would be globally expansive and store massive amounts of Earth’s carbon.
49
50
51 With regard to stability, magnesite occurring in freshwater playas has formed over 100s to 1000s
52
53 of years, suggesting a level of stability required for long-term storage of CO2.1 As is the case for
54
55
56
dolomite, the estimated precipitation rate for magnesite at Earth’s surface conditions11 may be
57
58
59
60
13
ACS Paragon Plus Environment
Crystal Growth & Design Page 14 of 29

1
2
3
sufficient to explain its occurrence in the geologic record, yet is exceedingly too slow to garner
4
5
6 interest as a pathway for permanent storage of anthropogenic CO2. Using high-density
7
8 carboxylated surfaces to accelerate magnesite formation at ambient temperature could lend itself
9
10
11
to securely storing CO2.
12
13 One approach to ex situ mineral carbonation involves the reaction of ultramafic rocks with
14
15 CO2 at high temperatures and pressures to form magnesite and silica.3 Although magnesite
16
17
18 formation occurs relatively rapidly at high temperature and pressure, carbonation rates may still
19
20 be limited by magnesite precipitation.29 Furthermore, the costs for pre-treatment, energy use,
21
22 and chemical inputs for ex situ mineral carbonation at high temperature (~$50-300/t CO2) greatly
23
24
25 exceed current carbon prices (e.g., California Carbon Allowance ≈ $13 US/t), severely limiting
26
27 the application of these technologies.3 Carbonation at ambient temperature and pressure
28
29
generally requires more favorable chemical conditions in terms of pH and Mg and dissolved
30
31
32 inorganic carbon concentrations for the precipitation of metastable hydrated Mg-carbonate
33
34 minerals than those required to reach magnesite saturation. For instance, nesquehonite
35
36
37 precipitates at near equilibrium conditions (e.g., SI ≈ 0) during aqueous carbonation of brucite
38
39 [Mg(OH)2], whereas magnesite remains highly supersaturated (e.g., SI ≈ 2.7).52 In our
40
41 experiments, precipitation occurred in solutions only slightly supersaturated with respect to
42
43
44 magnesite and without additional energy input. Moreover, microscopy of the final solids
45
46 indicates that the carboxylated spheres were not consumed during reaction, indicating that
47
48 further reaction is possible.
49
50
51 An alternative to silicate and hydroxide mineral feedstocks is to use evaporite deposits,
52
53 waste brines, produced waters from oil extraction, and a virtually unlimited supply of seawater as
54
55
56
feedstock for carbonate precipitation, thereby eliminating the need for mineral dissolution.1 The
57
58
59
60
14
ACS Paragon Plus Environment
Page 15 of 29 Crystal Growth & Design

1
2
3
abundance of readily available cation sources provides large CO2 sequestration capacity.1 For
4
5
6 instance, the oil industry generates an estimated 8 Gt/yr of produced water, and 10 Gt/yr of reject
7
8 brines with Mg concentrations ranging from 30 to 60 g/L and 0.5 to 20 g/L, respectively.53 In
9
10
11
comparison to seawater,54 the compositions of these waters are highly variable.55 Seawater is
12
13 supersaturated with respect to magnesite (SI ≈ 0.6), whereas hydrated Mg-carbonate minerals
14
15 such as hydromagnesite are unsaturated (SI ≈ -5.7), and therefore offer no carbon sequestration
16
17
18 capacity. On the other hand, equilibrium precipitation of magnesite from seawater would
19
20 sequester 2.0 g CO2/L if pH were maintained and seawater kept at equilibrium with atmospheric
21
22 CO2. Upscaling the processes of magnesite precipitation demonstrated in this study could prove
23
24
25 to be advantageous for ex situ carbonation strategies. The use of organic ligands for enhancing
26
27 silicate dissolution for mineral carbonation has been shown to have an inhibitory effect on
28
29
carbonate precipitation rates.56 Researchers attribute this effect to the formation of surface
30
31
32 complexes on the carbonate mineral or changing of the saturation state of the fluid.56-61
33
34 Geologic CO2 sequestration involves injection of CO2 into subsurface geologic formations
35
36
37 including saline aquifers as well as mafic and ultramafic bodies.62-65 Long-term (1000+ years)
38
39 storage of CO2 preferably relies on precipitation of carbonate minerals in the subsurface.
40
41 Magnesite has been produced by reacting synthetic forsterite with wet scCO2 at temperatures of
42
43
44 35 °C16, 47 and 50 °C66, conditions relevant to CO2 injection into geologic reservoirs. Magnesite
45
46 precipitation at 50 °C in scCO2 (90 bar) correlates with the equilibrium constants for
47
48 lignand−Mg2+ association reactions.47 Miller et al.47 suggest that ligand-promoted Mg2+
49
50
51 dehydration is likely enhanced in thin water films on mineral surfaces due to the highly ordered
52
53 nature of this water that has dielectric constants that are much lower than bulk water phase,
54
55
56
thereby enhancing ligand-H2O exchange reactions. The presence of organic ligands able to
57
58
59
60
15
ACS Paragon Plus Environment
Crystal Growth & Design Page 16 of 29

1
2
3
partially dehydrate Mg2+ ions in the subsurface at scCO2 injection sites may offer faster
4
5
6 precipitation rates for securely storing CO2. While upscaling our experimental systems remains
7
8 a challenge, the potential for removing kinetic barriers for magnesite precipitation has
9
10
11
implications for both ex situ and in situ carbonation processes at low temperature.
12
13
14
15 ASSOCIATED CONTENT
16
17
18 Supporting Information. Miller indices, lattice spacing and intensities for the aluminum
19
20 standard, additional scanning electron micrographs, as well as X-ray and electron diffraction
21
22 patterns of the aluminum standard and experimental precipitates.
23
24
25
26
27 AUTHOR INFORMATION
28
29
30
Corresponding Author
31
32 *E-mail: ipower@eoas.ubc.ca
33
34 ORDID
35
36
37 Ian Power: 0000-0003-2102-9315
38
39 Notes
40
41 The authors declare no competing financial interest.
42
43
44
45
46 ACKNOWLEDGEMENTS
47
48 Funding was provided by Natural Science and Engineering Research Council of Canada
49
50
51 (NSERC) Discovery grants to G.M. Dipple and M. Raudsepp. From The University of British
52
53 Columbia, we thank Gethin Owen at the Centre for High-Throughput Phenogenomics for
54
55
56
assistance with scanning electron microscopy and Bradford Ross at the Bioimaging Facility for
57
58
59
60
16
ACS Paragon Plus Environment
Page 17 of 29 Crystal Growth & Design

1
2
3
transmission electron microscopy. Special thanks to Todd Simpson at the Nanofabrication
4
5
6 Laboratory, Western University for thin section preparation.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
17
ACS Paragon Plus Environment
Crystal Growth & Design Page 18 of 29

1
2
3
REFERENCES
4
5
6 (1) Power, I. M.; Harrison, A. L.; Dipple, G. M.; Wilson, S. A.; Kelemen, P. B.; Hitch, M.;
7 Southam, G., Carbon mineralization: From natural analogues to engineered systems. In
8 Geochemistry of Geologic CO2 Sequestration, DePaolo, D. J.; Cole, D. R.; Navrotsky, A.;
9 Bourg, I. C., Eds. The Mineralogical Society of America: Chantilly, Virginia, U.S.A., 2013; Vol.
10
11
77, pp 305-360.
12 (2) De Yoreo, J. J.; Waychunas, G. A.; Jun, Y. S.; Fernandez-Martinez, A., In situ
13 Investigations of Carbonate Nucleation on Mineral and Organic Surfaces. In Geochemistry of
14 Geologic CO2 Sequestration, DePaolo, D. J.; Cole, D. R.; Navrotsky, A.; Bourg, I. C., Eds.
15 Mineralogical Soc Amer: Chantilly, 2013; Vol. 77, pp 229-257.
16
(3) Sanna, A.; Uibu, M.; Caramanna, G.; Kuusik, R.; Maroto-Valer, M. M., A review of
17
18 mineral carbonation technologies to sequester CO2. Chemical Society Reviews 2014, 43, (23),
19 8049-8080.
20 (4) Harrison, A. L.; Dipple, G. M.; Power, I. M.; Mayer, K. U., Influence of surface
21 passivation and water content on mineral reactions in unsaturated porous media: Implications for
22 brucite carbonation and CO2 sequestration. Geochim. Cosmochim. Acta 2015, 148, 477-495.
23
24
(5) Wilson, S. A.; Harrison, A. L.; Dipple, G. M.; Power, I. M.; Barker, S. L. L.; Mayer, K.
25 U.; Fallon, S. J.; Raudsepp, M.; Southam, G., Offsetting of CO2 emissions by air capture in mine
26 tailings at the Mount Keith Nickel Mine, Western Australia: Rates, controls and prospects for
27 carbon neutral mining. Int. J. Greenh. Gas Control 2014, 25, 121-140.
28 (6) McCutcheon, J.; Power, I. M.; Harrison, A. L.; Dipple, G. M.; Southam, G., A
29
Greenhouse-Scale Photosynthetic Microbial Bioreactor for Carbon Sequestration in Magnesium
30
31 Carbonate Minerals. Environ. Sci. Technol. 2014, 48, (16), 9142-9151.
32 (7) Beinlich, A.; Austrheim, H., In situ sequestration of atmospheric CO2 at low temperature
33 and surface cracking of serpentinized peridotite in mine shafts. Chem. Geol. 2012, 332, 32-44.
34 (8) Wilson, S. A.; Barker, S. L. L.; Dipple, G. M.; Atudorei, V., Isotopic disequilibrium
35 during uptake of atmospheric CO2 into mine process waters: Implications for CO2 sequestration.
36
37 Environ. Sci. Technol. 2010, 44, 9522-9529.
38 (9) Montes-Hernandez, G.; Renard, F., Time-resolved in situ raman spectroscopy of the
39 nucleation and growth of siderite, magnesite, and calcite and their precursors. Cryst. Growth
40 Des. 2016, 16, (12), 7218-7230.
41 (10) Königsberger, E.; Königsberger, L.; Gamsjager, H., Low-temperature thermodynamic
42
43
model for the system Na2CO3-MgCO3-CaCO3-H2O. Geochim. Cosmochim. Acta 1999, 63, 3105-
44 3119.
45 (11) Saldi, G. D.; Jordan, G.; Schott, J.; Oelkers, E. H., Magnesite growth rates as a function
46 of temperature and saturation state. Geochim. Cosmochim. Acta 2009, 73, (19), 5646-5657.
47 (12) Schott, J.; Pokrovsky, O. S.; Oelkers, E. H., The Link Between Mineral
48 Dissolution/Precipitation Kinetics and Solution Chemistry. In Thermodynamics and Kinetics of
49
50 Water-Rock Interaction, Oelkers, E. H.; Schott, J., Eds. Mineralogical Soc Amer: Chantilly,
51 2009; Vol. 70, pp 207-258.
52 (13) Gregg, J. M.; Bish, D. L.; Kaczmarek, S. E.; Machel, H. G., Mineralogy, nucleation and
53 growth of dolomite in the laboratory and sedimentary environment: A review. Sedimentology
54 2015, 62, (6), 1749-1769.
55
56
57
58
59
60
18
ACS Paragon Plus Environment
Page 19 of 29 Crystal Growth & Design

1
2
3
(14) Qafoku, O.; Dixon, D. A.; Rosso, K. M.; Schaef, H. T.; Bowden, M. E.; Arey, B. W.;
4
5 Felmy, A. R., Dynamics of Magnesite Formation at Low Temperature and High pCO2 in
6 Aqueous Solution. Environ. Sci. Technol. 2015, 49, (17), 10736-10744.
7 (15) Qafoku, O.; Hu, J.; Hess, N. J.; Hu, M. Y.; Ilton, E. S.; Feng, J.; Arey, B. W.; Felmy, A.
8 R., Formation of submicron magnesite during reaction of natural forsterite in H2O-saturated
9 supercritical CO2. Geochim. Cosmochim. Acta 2014, 134, 197-209.
10
11
(16) Felmy, A. R.; Qafoku, O.; Arey, B. W.; Hu, J. Z.; Hu, M.; Schaef, H. T.; Ilton, E. S.;
12 Hess, N. J.; Pearce, C. I.; Feng, J.; Rosso, K. M., Reaction of water-saturated supercritical CO2
13 with forsterite: Evidence for magnesite formation at low temperatures. Geochim. Cosmochim.
14 Acta 2012, 91, 271-282.
15 (17) Felmy, A. R.; Qafoku, O.; Arey, B. W.; Kovarik, L.; Liu, J.; Perea, D.; Ilton, E. S.,
16
Enhancing magnesite formation at low temperature and high CO2 pressure: The impact of seed
17
18 crystals and minor components. Chem. Geol. 2015, 395, 119-125.
19 (18) Hänchen, M.; Prigiobbe, V.; Baciocchi, R.; Mazzotti, M., Precipitation in the Mg-
20 carbonate system - effects of temperature and CO2 pressure. Chemical Engineering Science
21 2008, 63, (4), 1012-1028.
22 (19) Land, L. S., Failure to precipitate dolomite at 25 °C from dilute solution despite 1000-
23
24
fold oversaturation after 32 years. Aquatic Geochemistry 1998, 4, 361-368.
25 (20) Vasconcelos, C.; McKenzie, J. A.; Bernasconi, S.; Grujic, D.; Tien, A. J., Microbial
26 mediation as a possible mechanism for natural dolomite formation at low temperatures. Nature
27 1995, 377, (6546), 220-222.
28 (21) Kenward, P. A.; Fowle, D. A.; Goldstein, R. H.; Ueshima, M.; Gonzalez, L. A.; Roberts,
29
J. A., Ordered low-temperature dolomite mediated by carboxyl-group density of microbial cell
30
31 walls. AAPG Bull. 2013, 97, (11), 2113-2125.
32 (22) Roberts, J. A.; Kenward, P. A.; Fowle, D. A.; Goldstein, R. H.; Gonzalez, L. A.; Moore,
33 D. S., Surface chemistry allows for abiotic precipitation of dolomite at low temperature. Proc.
34 Natl. Acad. Sci. U. S. A. 2013, 110, (36), 14540-14545.
35 (23) Zhang, F. F.; Xu, H. F.; Konishi, H.; Shelobolina, E. S.; Roden, E. E., Polysaccharide-
36
37 catalyzed nucleation and growth of disordered dolomite: A potential precursor of sedimentary
38 dolomite. American Mineralogist 2012, 97, (4), 556-567.
39 (24) Parkhurst, D. L.; Appelo, C. A. J., Description of Input and Examples for PHREEQC
40 Version 3 — A Computer Program for Speciation, Batch-Reaction, One-Dimensional Transport,
41 and Inverse Geochemical Calculations. In U.S. Geological Survey Techniques and Methods,
42
43
Denver, Colorado, 2013; Vol. book 6, p 497.
44 (25) Bruker AXS, DIFFRACplus EVA 14 Release 2008; Karlsruhe, Germany, 2008.
45 (26) Power, I. M.; Wilson, S. A.; Harrison, A. L.; Dipple, G. M.; McCutcheon, J.; Southam,
46 G.; Kenward, P. A., A depositional model for hydromagnesite–magnesite playas near Atlin,
47 British Columbia, Canada. Sedimentology 2014, 61, 1701-1733.
48 (27) Falk, E. S.; Kelemen, P. B., Geochemistry and petrology of listvenite in the Samail
49
50 ophiolite, Sultanate of Oman: Complete carbonation of peridotite during ophiolite emplacement.
51 Geochim. Cosmochim. Acta 2015, 160, 70-90.
52 (28) Swanson, E. J.; Fricker, K. J.; Sun, M.; Park, A. H. A., Directed precipitation of hydrated
53 and anhydrous magnesium carbonates for carbon storage. Phys. Chem. Chem. Phys. 2014, 16,
54 (42), 23440-23450.
55
56
(29) Saldi, G. D.; Schott, J.; Pokrovsky, O. S.; Gautier, Q.; Oelkers, E. H., An experimental
57 study of magnesite precipitation rates at neutral to alkaline conditions and 100-200 degrees C as
58
59
60
19
ACS Paragon Plus Environment
Crystal Growth & Design Page 20 of 29

1
2
3
a function of pH, aqueous solution composition and chemical affinity. Geochim. Cosmochim.
4
5 Acta 2012, 83, 93-109.
6 (30) Power, I. M.; Wilson, S. A.; Thom, J. M.; Dipple, G. M.; Gabites, J. E.; Southam, G., The
7 hydromagnesite playas of Atlin, British Columbia, Canada: A biogeochemical model for CO2
8 sequestration. Chem. Geol. 2009, 260, (3-4), 286-300.
9 (31) Power, I. M.; Wilson, S. A.; Thom, J. M.; Dipple, G. M.; Southam, G., Biologically
10
11
induced mineralization of dypingite by cyanobacteria from an alkaline wetland near Atlin,
12 British Columbia, Canada. Geochemical Transactions 2007, 8, 13.
13 (32) Power, I. M.; Harrison, A. L.; Dipple, G. M., Accelerating Mineral Carbonation Using
14 Carbonic Anhydrase. Environ. Sci. Technol. 2016, 50, (5), 2610-2618.
15 (33) Harrison, A. L.; Dipple, G. M.; Power, I. M.; Mayer, K. U., The impact of evolving
16
mineral-water-gas interfacial areas on mineral-fluid reaction rates in unsaturated porous media.
17
18 Chem. Geol. 2016, 421, 65-80.
19 (34) Ming, D. W.; Franklin, W. T., Synthesis and characterization of lansfordite and
20 nesquehonite. Soil Science Society of America Journal 1985, 49, (5), 1303-1308.
21 (35) Wilson, S. A.; Dipple, G. M.; Power, I. M.; Barker, S. L. L.; Fallon, S. J.; Southam, G.,
22 Subarctic Weathering of Mineral Wastes Provides a Sink for Atmospheric CO2. Environ. Sci.
23
24
Technol. 2011, 45, (18), 7727-7736.
25 (36) Di Lorenzo, F.; Rodríguez-Galán, R. M.; Prieto, M., Kinetics of the solvent-mediated
26 transformation of hydromagnesite into magnesite at different temperatures. Mineral. Mag. 2014,
27 78, (6), 1363-1372.
28 (37) Jiao, D.; King, C.; Grossfield, A.; Darden, T. A.; Ren, P. Y., Simulation of Ca2+ and Mg2+
29
solvation using polarizable atomic multipole potential. Journal of Physical Chemistry B 2006,
30
31 110, (37), 18553-18559.
32 (38) Kluge, S.; Weston, J., Can a hydroxide ligand trigger a change in the coordination
33 number of magnesium ions in biological systems. Biochemistry 2005, 44, 4877-4885.
34 (39) Di Tommaso, D.; de Leeuw, N. H., Structure and dynamics of the hydrated magnesium
35 ion and of the solvated magnesium carbonates: insights from first principles simulations. Phys.
36
37 Chem. Chem. Phys. 2010, 12, (4), 894-901.
38 (40) Markham, G. D.; Glusker, J. P.; Bock, C. W., The arrangement of first- and second-
39 sphere water molecules in divalent magnesium complexes: Results from molecular orbital and
40 density functional theory and from structural crystallography. Journal of Physical Chemistry B
41 2002, 106, (19), 5118-5134.
42
43
(41) Bock, C. W.; Kaufman, A.; Glusker, J. P., Coordination of water to magnesium cations.
44 Inorg. Chem. 1994, 33, (3), 419-427.
45 (42) Dudev, T.; Cowan, J. A.; Lim, C., Competitive binding in magnesium coordination
46 chemistry: Water versus ligands of biological interest. J. Am. Chem. Soc. 1999, 121, 7665-7673.
47 (43) Wang, D.; Wallace, A. F.; De Yoreo, J. J.; Dove, P. M., Carboxylated molecules regulate
48 magnesium content of amorphous calcium carbonates during calcification. Proc. Natl. Acad. Sci.
49
50 U. S. A. 2009, 106, (51), 21511-21516.
51 (44) Schultze-Lam, S.; Fortin, D.; Davis, B. S.; Beveridge, T. J., Mineralization of bacterial
52 surfaces. Chem. Geol. 1996, 132, (1-4), 171-181.
53 (45) Sikiric, M. D.; Fueredi-Milhofer, H., The influence of surface active molecules on the
54 crystallization of biominerals in solution. Advances in Colloid and Interface Science 2006, 128,
55
56
135-158.
57
58
59
60
20
ACS Paragon Plus Environment
Page 21 of 29 Crystal Growth & Design

1
2
3
(46) Sanz-Montero, M. E.; Rodríguez-Aranda, J. P., Magnesite formation by microbial
4
5 activity: Evidence from a Miocene hypersaline lake. Sediment. Geol. 2012, 263, 6-15.
6 (47) Miller, Q. R. S.; Kaszuba, J. P.; Schaef, H. T.; Bowden, M. E.; McGrail, B. P., Impacts of
7 Organic Ligands on Forsterite Reactivity in Supercritical CO2 Fluids. Environ. Sci. Technol.
8 2015, 49, (7), 4724-4734.
9 (48) de Leeuw, N. H.; Parker, S. C., Surface-water interactions in the dolomite problem. Phys.
10
11
Chem. Chem. Phys. 2001, 3, (15), 3217-3221.
12 (49) Xu, J.; Yan, C.; Zhang, F. F.; Konishi, H.; Xu, H. F.; Teng, H. H., Testing the cation-
13 hydration effect on the crystallization of Ca-Mg-CO3 systems. Proc. Natl. Acad. Sci. U. S. A.
14 2013, 110, (44), 17750-17755.
15 (50) Melezhik, V. A.; Fallick, A. E.; Medvedev, P. V.; Makarikhin, V. V., Palaeoproterozoic
16
magnesite: lithological and isotopic evidence for playa/sabkha environments. Sedimentology
17
18 2001, 48, (2), 379-397.
19 (51) Wright, D. T.; Wacey, D., Precipitation of dolomite using sulphate-reducing bacteria
20 from the Coorong Region, South Australia: significance and implications. Sedimentology 2005,
21 52, 987-1008.
22 (52) Harrison, A. L.; Power, I. M.; Dipple, G. M., Accelerated Carbonation of Brucite in Mine
23
24
Tailings for Carbon Sequestration. Environ. Sci. Technol. 2013, 47, (1), 126-134.
25 (53) Mignardi, S.; De Vito, C.; Ferrini, V.; Martin, R. F., The efficiency of CO2 sequestration
26 via carbonate mineralization with simulated wastewaters of high salinity. J. Hazard. Mater.
27 2011, 191, (1-3), 49-55.
28 (54) Nordstrom, D. K.; Plummer, L. N.; Wigley, T. M. L.; Wolery, T. J.; Ball, J. W.; Jenne, E.
29
A.; Bassett, R. L.; Crerar, D. A.; T.M., F.; Fritz, B.; Hoffman, M.; Holdren, G. R., Jr. ; Lafon, G.
30
31 M.; Mattigod, S. V.; McDuff, R. E.; Morel, F.; Reddy, M. M.; Sposito, G.; Thrailkill, J., A
32 comparison of computerized chemical models for equilibrium calculations in aqueous systems.
33 In Chemical Modeling in aqueous systems, speciation, sorption, solubility, and kinetics, Jenne,
34 E. A., Ed. American Chemical Society: 1979; pp p. 857-892.
35 (55) Blondes, M. S.; Gans, K. D.; Rowan, E. L.; Thordsen, J. J.; Reidy, M. E.; Engle, M. A.;
36
37 Kharaka, Y. K.; Thomas, B., U.S. Geological Survey National Produced Waters Geochemical
38 Database v2.2. 2017.
39 (56) Gautier, Q.; Berninger, U.-N.; Schott, J.; Jordan, G., Influence of organic ligands on
40 magnesite growth: A hydrothermal atomic force microscopy study. Geochim. Cosmochim. Acta
41 2015, 155, 68-85.
42
43
(57) Bonfils, B.; Julcour-Lebigue, C.; Guyot, F.; Bodénan, F.; Chiquet, P.; Bourgeois, F.,
44 Comprehensive analysis of direct aqueous mineral carbonation using dissolution enhancing
45 organic additives. Int. J. Greenh. Gas Control 2012, 9, 334-346.
46 (58) Gautier, Q.; Bénézeth, P.; Schott, J., Magnesite growth inhibition by organic ligands: An
47 experimental study at 100, 120 and 146 degrees °C. Geochim. Cosmochim. Acta 2016, 181, 101-
48 125.
49
50 (59) Reddy, M. M.; Hoch, A. R., Calcite crystal growth rate inhibition by polycarboxylic
51 acids. Journal of Colloid and Interface Science 2001, 235, (2), 365-370.
52 (60) Westin, K.; Rasmuson, A. C., Crystal growth of aragonite and calcite in presence of citric
53 acid, DTPA, EDTA and pyromellitic acid. Journal of Colloid and Interface Science 2005, 282,
54 (2), 359-369.
55
56
57
58
59
60
21
ACS Paragon Plus Environment
Crystal Growth & Design Page 22 of 29

1
2
3
(61) Ganor, J.; Reznik, I. J.; Rosenberg, Y. O., Organics in Water-Rock Interactions. In
4
5 Thermodynamics and Kinetics of Water-Rock Interaction, Oelkers, E. H.; Schott, J., Eds.
6 Mineralogical Soc Amer: Chantilly, 2009; Vol. 70, pp 259-369.
7 (62) Thom, J.; Dipple, G. M. In Serpentine dissolution kinetics and rates of CO2
8 sequestration, American Geophysical Union Fall Meeting Supplement, San Francisco,
9 California. Abstract V53D-1785, 2006; San Francisco, California. Abstract V53D-1785, 2006.
10
11
(63) Power, I. M.; Wilson, S. A.; Dipple, G. M., Serpentinite carbonation for CO2
12 sequestration. Elements 2013, 9, 115-121.
13 (64) Cipolli, F.; Gambardella, B.; Marini, L.; Ottonello, G.; Zuccolini, M. V., Geochemistry of
14 high-pH waters from serpentinites of the Gruppo di Voltri (Genova, Italy) and reaction path
15 modeling of CO2 sequestration in serpentinite aquifers. Appl. Geochem. 2004, 19, (5), 787-802.
16
(65) Matter, J. M.; Stute, M.; Snaebjörnsdottir, S. O.; Oelkers, E. H.; Gislason, S. R.;
17
18 Aradottir, E. S.; Sigfusson, B.; Gunnarsson, I.; Sigurdardottir, H.; Gunnlaugsson, E.; Axelsson,
19 G.; Alfredsson, H. A.; Wolff-Boenisch, D.; Mesfin, K.; Taya, D. F. D. l. R.; Hall, J.; Dideriksen,
20 K.; Broecker, W. S., Rapid carbon mineralization for permanent disposal of anthropogenic
21 carbon dioxide emissions. Science 2016, 352, (6291), 1312-1314.
22 (66) Schaef, H. T.; McGrail, B. P.; Loring, J. L.; Bowden, M. E.; Arey, B. W.; Rosso, K. M.,
23
24
Forsterite [Mg2SiO4)] Carbonation in Wet Supercritical CO2: An in Situ High-Pressure X-ray
25 Diffraction Study. Environ. Sci. Technol. 2013, 47, (1), 174-181.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
22
ACS Paragon Plus Environment
Page 23 of 29 Crystal Growth & Design

1
2
3
Figure captions
4
5
6 Fig. 1. Representative scanning electron micrographs of polystyrene microspheres and
7
8 associated mineral precipitates after 60 days of reaction. (A) Low carboxyl site density spheres
9
10
11
with unidentified mineral precipitate. (B) Mineral precipitate on high carboxyl site density
12
13 sphere showing step growth texture. (C) Numerous mineral precipitates (arrows) held in an
14
15 organic matrix attached to a sphere (upper right). (D) Mineral precipitates on filter paper
16
17
18 exhibiting attachment and aggregation. (E) For comparison, naturally occurring magnesite from
19
20 a hydromagnesite-magnesite playa. (F) Representative energy dispersive spectrum of mineral
21
22 precipitate (G) revealing the chemical composition of Mg, C and O (Os is from coating). Scale
23
24
25 bars are 500 nm.
26
27
28
29
Fig. 2. Electron micrographs and SAED pattern of experimental precipitate. (A) Mineral
30
31
32 precipitate that was thin sectioned using FIB milling (B). (C) Representative transmission
33
34 electron micrograph of a thin section (~100 nm thickness) of a mineral grain obtained by FIB
35
36
37 milling. Note the outlines of rhombohedron, characteristic of magnesite (arrows). (D) An
38
39 SAED pattern obtained from the mineral grain section viewing down the [3 0 1] plane.
40
41 Reflections (h k l) are labeled with the measured lattice spacings that closely match those of
42
43
44 magnesite (Table 1). Scale bars are 500 nm when not labeled.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
23
ACS Paragon Plus Environment
Crystal Growth & Design Page 24 of 29

1
2
3
Table 1. List of reflections and their associated lattice spacings for magnesite as compared
4
5
6 to those measured for the unknown mineral precipitate (Fig. 2).
7
8 Reflection Magnesite Measured
9 hkl lattice spacing (Å) lattice spacing (Å)
10 123 2.1022 2.10
11 113 2.1022 2.11
12 216 1.6999 1.70
13 030 1.3374 1.34
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
24
ACS Paragon Plus Environment
Page 25 of 29 Crystal Growth & Design

1
2
3
Figure 1
4
5
6 A B
7 sphere
8
9
10
11
12
13
14
15 sphere
16
17 C D
18 sphere
19
organic
20
matrix
21
22
23
24
25
26
27
28
E F Mg G
29
Intensity (counts in 1000s)

8
30
31
6 O
32
mgs
33 4
34
35 2 C
36 Os
37 hmg 0
0 1 2 3 4 5 6 7 8 9 10
38 Energy (keV)
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
25
ACS Paragon Plus Environment
Crystal Growth & Design Page 26 of 29

1
2
3
Figure 2
4
5
6 A B
7
8
9
10
11
12
13
14
15
5 µm
16
17 C D [3 0 1]
18 22 11 66
19
20 11 2 33

21 00 33 00
1.34
Å
22 11 11 33
23 Å
2.11
24

2 .1
Å


25 11 11 33

1.70
00 3 0
26
27 11 22 33
28 22 1 6
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
26
ACS Paragon Plus Environment
Page 27 of 29 Crystal Growth & Design

1
2
3
For Table of Contents Use Only
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Magnesite (MgCO3) is one of the most stable sinks for carbon dioxide (CO2) and is of
20
21 great interest for long-term carbon storage. Here, we document the precipitation of magnesite in
22
23 batch reactors at room temperature with the aid of carboxylated polystyrene microspheres.
24
25
26 Research findings may lead to novel and economically viable strategies for permanent carbon
27
28 storage.
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
27
ACS Paragon Plus Environment
A B & Design
Crystal Growth Page 28 of 29
sphere

1
2
3
4
5
6
7 sphere
8
C
9 D
sphere
10
11 organic
12 matrix
13
14
15
16
17
18
19
E
20 F Mg G
Intensity (counts in 1000s)

21 8
22
23 6 O
24 mgs
25 4

26
27 2 C
28 ACS Paragon Plus Environment
Os
hmg 0
29 0 1 2 3 4 5 6 7 8 9 10
30 Energy (keV)
A
Page 29 of 29 B & Design
Crystal Growth

1
2
3
4
5
6
7 5 μm
8
9C D [3 0 1]
10 22 11 66

11 11 22 33
12 1.34
00 33 00 Å
13 11 11 33
14 Å
2.11
15

2.1
16
Å
11 11 33


17 1.70 00 33 00
18 ACS Paragon Plus Environment
11 22 33
19
22 11 66
20

You might also like