10.08.2013 Views

Author's personal copy - University of Brighton Repository

Author's personal copy - University of Brighton Repository

Author's personal copy - University of Brighton Repository

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

This article appeared in a journal published by Elsevier. The attached<br />

<strong>copy</strong> is furnished to the author for internal non-commercial research<br />

and education use, including for instruction at the authors institution<br />

and sharing with colleagues.<br />

Other uses, including reproduction and distribution, or selling or<br />

licensing copies, or posting to <strong>personal</strong>, institutional or third party<br />

websites are prohibited.<br />

In most cases authors are permitted to post their version <strong>of</strong> the<br />

article (e.g. in Word or Tex form) to their <strong>personal</strong> website or<br />

institutional repository. Authors requiring further information<br />

regarding Elsevier’s archiving and manuscript policies are<br />

encouraged to visit:<br />

http://www.elsevier.com/<strong>copy</strong>right


<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO 2–H 2O during<br />

CO 2 degassing <strong>of</strong> Mg-bearing solutions<br />

Laurence Hopkinson a,⇑ , Petra Kristova b,1 , Ken Rutt b,2 , Gordon Cressey c,3<br />

a School <strong>of</strong> Environment and Technology, <strong>University</strong> <strong>of</strong> <strong>Brighton</strong>, Cockcr<strong>of</strong>t Building, Lewes Road, <strong>Brighton</strong> BN2 4GJ, United Kingdom<br />

b School <strong>of</strong> Pharmacy and Biomolecular Sciences, <strong>University</strong> <strong>of</strong> <strong>Brighton</strong>, Cockcr<strong>of</strong>t Building, Lewes Road, <strong>Brighton</strong> BN2 4GJ, United Kingdom<br />

c Department <strong>of</strong> Mineralogy, Natural History Museum, Cromwell Road, London SW7 5BD, United Kingdom<br />

Abstract<br />

Received 5 January 2011; accepted in revised form 12 October 2011; available online 18 October 2011<br />

This study documents the paragenesis <strong>of</strong> magnesium carbonates formed during degassing <strong>of</strong> CO2 from a 0.15 M Mg 2+<br />

aqueous solution. The starting solutions were prepared by CO2 sparging <strong>of</strong> a brucite suspension at 25 °C for 19 h, followed<br />

by rapid heating to 58 °C. One experiment was performed in an agitated environment, promoted by sonication. In the second,<br />

CO2 degassing was exclusively thermally-driven (static environment). Electric conductance, pH, and temperature <strong>of</strong> the experimental<br />

solutions were measured, whereas Mg 2+ was determined by atomic absorption spectros<strong>copy</strong>. Precipitates were analysed<br />

by X-ray diffraction, Fourier transform (FT) mid-infrared, FT-Raman, and scanning electron micros<strong>copy</strong>.<br />

Hydromagnesite [Mg5(CO3)4(OH)2 4H2O] precipitated at 25 °C was followed by nesquehonite [Mg(HCO3,OH) 2H2O]<br />

upon heating to 58 °C. The yield <strong>of</strong> the latter mineral was greater in the agitated solution. After 120 min, accelerated CO2<br />

degassing resulted in the loss <strong>of</strong> nesquehonite at the expense <strong>of</strong> an assemblage consisting <strong>of</strong> an unnamed mineral phase:<br />

[Mg5(CO3)4(OH)2 8H2O] and hydromagnesite. After 240 min, dypingite [Mg5(CO3)4(OH)2 5H2O (or less H2O)] appears with<br />

hydromagnesite. The unnamed mineral shows greater disorder than dypingite, which in turn shows greater disorder than hydromagnesite.<br />

In the static environment, there is no evidence for nesquehonite loss or the generation <strong>of</strong> [Mg5(CO3)4(OH)2 X-<br />

H 2O] phases over the same timeframe. Hence, results indicate that the transformation <strong>of</strong> nesquehonite to hydromagnesite<br />

displays mixed diffusion and reaction-limited control and proceeds through the production <strong>of</strong> metastable intermediates,<br />

and is interpreted according to the Ostwald step rule. Nevertheless, variations in the chemistry <strong>of</strong> nesquehonite, combined<br />

with the established tendency <strong>of</strong> the mineral to desiccate, implies that its transformation to hydromagnesite is unlikely to follow<br />

a single simple sequential reaction pathway.<br />

Ó 2011 Elsevier Ltd. All rights reserved.<br />

1. INTRODUCTION<br />

Available online at www.sciencedirect.com<br />

Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

For several important groups <strong>of</strong> minerals, including<br />

carbonates, mineral paragenesis frequently results in<br />

⇑ Corresponding author. Tel.: +44 (0) 1273 642239; fax: +44 (0)<br />

1273 642285.<br />

E-mail addresses: l.hopkinson@brighton.ac.uk (L. Hopkinson),<br />

p.kristova@brighton.ac.uk (P. Kristova), k.rutt@brighton.ac.uk<br />

(K. Rutt), g.cressey@nhm.ac.uk (G. Cressey).<br />

1<br />

Tel.: +44 (0) 1273 642065; fax: +44 (0) 1273 642285.<br />

2<br />

Tel.: +44 (0) 1273 642076; fax: +44 (0) 1273 642285.<br />

3<br />

Tel.: +44 (0) 20 7942 5711; fax: +44 (0) 20 7942 5537.<br />

0016-7037/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.gca.2011.10.023<br />

www.elsevier.com/locate/gca<br />

assemblages that do not conform to predictions based on<br />

equilibrium thermodynamics (e.g., Morse and Casey,<br />

1988). In the system MgO–CO2–H2O at 0–60 °C, the stable<br />

phases in equilibrium with aqueous solutions containing<br />

Mg ions and carbon dioxide are brucite [Mg(OH) 2] at very<br />

low partial pressures <strong>of</strong> CO2, and magnesite [MgCO3], the<br />

most stable magnesium carbonate under Earth surface conditions<br />

(e.g., Königsberger et al., 1999; Hänchen et al.,<br />

2008). Nevertheless, Mg-carbonate precipitation is strongly<br />

kinetically-controlled (e.g., Königsberger et al., 1999). Inhibition<br />

<strong>of</strong> magnesite formation has been ascribed to the high<br />

hydration energy <strong>of</strong> Mg 2+ (e.g., Hänchen et al., 2008).<br />

Instead <strong>of</strong> magnesite, a variety <strong>of</strong> magnesium hydrate and


asic carbonates (i.e. containing hydroxyl groups) form, <strong>of</strong><br />

which hydromagnesite [Mg5(CO3)4(OH)2 4H2O (or 5H2O)]<br />

is the only stable mineral at atmospheric CO2 pressures<br />

within the temperature interval typical <strong>of</strong> most surface environments<br />

(Langmuir, 1965). Nevertheless, a number <strong>of</strong><br />

hydromagnesite-like phases, which differ in unit cell parameters<br />

and in some cases, numbers <strong>of</strong> waters <strong>of</strong> crystallization,<br />

also exist (e.g., Raade, 1970; Davies and Bubela,<br />

1973; Canterford et al., 1984). The relationship between<br />

these mineral phases and hydromagnesite as well as the<br />

causative factors behind their genesis in preference to hydromagnesite<br />

are uncertain, although it seems that the existence<br />

<strong>of</strong> hydromagnesite-like phases is, at least in part,<br />

accountable for discrepancies in solubility data for hydromagnesite<br />

reported in the literature (Königsberger et al.,<br />

1999).<br />

The mineral nesquehonite, reported as either<br />

[MgCO 3 3H 2O] or [Mg(HCO 3, OH) 2H 2O], forms at ca<br />

10–50 °C and higher than atmospheric PCO2 (e.g., Wells,<br />

1915; Beck, 1950; Kazakov et al., 1959; White, 1971; Stephan<br />

and MacGillavry, 1972; Coleyshaw et al., 2003; Hales<br />

et al., 2008; Cheng and Li, 2010). The mineral is unstable at<br />

near surface ambient conditions. In experimental settings,<br />

nesquehonite transforms to hydromagnesite [N ! HM] in<br />

an aqueous medium at 52–65 °C, depending on the reaction<br />

time (e.g., Davies and Bubela, 1973; Zhang et al., 2006).<br />

This is achieved through one or more dissolution–precipitation<br />

step(s), giving rise to a variety <strong>of</strong> short-lived metastable<br />

intermediates, which in some experimental settings include<br />

hydromagnesite-like phases (Davies and Bubela, 1973;<br />

Hopkinson et al., 2008). Moreover, the direct transformation<br />

<strong>of</strong> nesquehonite to hydromagnesite has also been suggested<br />

(Hao and Du, 2009). Hence, nesquehonite acts as a<br />

precursor for at least some hydromagnesite and, in some<br />

experiments, hydromagnesite-like phases are also generated<br />

during the [N ! HM] transition, the stability <strong>of</strong> the minerals<br />

influenced both by temperature and partial pressure <strong>of</strong><br />

CO2 (Königsberger et al., 1999). It follows that understanding<br />

the crystallization behaviour <strong>of</strong> hydrated magnesium<br />

carbonates continues to be a challenge and that factors<br />

determining products <strong>of</strong> the [N ! HM] transition are<br />

poorly characterized. To these ends, this study documents<br />

the nature <strong>of</strong> the metastable intermediates, generated in<br />

the course <strong>of</strong> the [N ! HM] transition, during open system<br />

degassing <strong>of</strong> CO2 from heated aqueous solutions.<br />

2. HYDRATED AND BASIC MAGNESIUM<br />

CARBONATES<br />

A number <strong>of</strong> magnesium carbonates have been identified<br />

in the system MgO–CO2–H2O (Fig. 1). Phases <strong>of</strong> relevance<br />

to this study are [Mg5(CO3)4(OH)2 XH2O] and [MgCO3 X-<br />

H 2O] group minerals. The [Mg 5(CO 3) 4(OH) 2] group have<br />

four to eleven H 2O molecules: heavy (5H 2O) and light<br />

(4H2O) hydromagnesite (e.g., Canterford et al., 1984;<br />

Botha and Strydom, 2001), dypingite [Mg5(CO3)4(OH)2<br />

5H 2O (or 6H 2O)] (Raade, 1970), giorgiosite [Mg 5(CO 3) 4<br />

(OH)2 6H2O] (Raade, 1970; Friedel, 1975), an unnamed<br />

mineral [Mg5(CO3)4(OH)2 8H2O] (Suzuki and Ito, 1973),<br />

and protohydromagnesite [Mg 5(CO 3) 4(OH) 2 11H 2O]<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

2 L. Hopkinson et al. / Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

(Davies and Bubela, 1973; Davies et al., 1977; Canterford<br />

et al., 1984). In the MgO–CO2–H2O system, these<br />

[Mg5(CO3)4(OH)2 XH2O] mineral phases fall on a line that<br />

connects protohydromagnesite with hydromagnesite<br />

(Fig. 1).<br />

Hydromagnesite is by far the most common naturallyoccurring<br />

mineral in the group. It occurs as a weathering<br />

product <strong>of</strong> magnesium-rich rocks, and as deposits in lakes,<br />

alkaline wetlands and evaporitic basins (e.g., Stamatakis,<br />

1995; Russell et al., 1999; Power et al., 2009; Last et al.,<br />

2010). Accounts <strong>of</strong> dypingite are far less widespread,<br />

although it has been described in similar settings (e.g., Raade,<br />

1970), frequently in association with hydromagnesite (e.g.,<br />

Power et al., 2007; Last et al., 2010). Rarer still, giorgiosite<br />

and the unnamed mineral are described in association with<br />

weathered volcanics (Suzuki and Ito, 1973; Friedel, 1975).<br />

Dypingite and protohydromagnesite have been synthesised<br />

individually during the [N ! HM] transition in separate<br />

studies (Davies and Bubela, 1973; Hopkinson et al., 2008).<br />

Protohydromagnesite has not been identified in nature.<br />

The vibrational spectra <strong>of</strong> the [Mg5(CO3) 4(OH) 2 XH2O] phases are undocumented except for hydromagnesite and<br />

dypingite. The latter mineral shows infrared-active internal<br />

modes <strong>of</strong> the [CO 2<br />

3 ] anion that are indistinguishable from<br />

those <strong>of</strong> hydromagnesite (e.g., Raade, 1970; Canterford<br />

et al., 1984). This implies that the two minerals share very<br />

similar or identical short-range order, yet, distinct longrange<br />

crystal order. More recent spectroscopic studies <strong>of</strong><br />

dypingite show that variations in the intensity <strong>of</strong> water<br />

stretching bands occurs, indicating that differences in the<br />

chemical formula exist, and that [CO 2<br />

3 ] units are variably<br />

distorted (Frost et al., 2008, 2009). Heating dypingite results<br />

in its conversion to hydromagnesite (Raade, 1970;<br />

Botha and Strydom, 2001).<br />

Known naturally-occurring [MgCO3 XH2O] minerals<br />

are: landsfordite, nesquehonite and barringtonite (Fig. 1).<br />

Landsfordite and nesquehonite are typically found in caves,<br />

or associated with weathered surfaces <strong>of</strong> magnesium-rich<br />

rocks, where temperatures are low and CO2 partial pressures<br />

are <strong>of</strong>ten more than ten times atmospheric (Langmuir,<br />

1965). At a CO2 partial pressure <strong>of</strong> 1 atmosphere, landsfordite<br />

transforms to nesquehonite at 10 ± 2 °C. Replacement<br />

is spontaneous and irreversible (Hill et al., 1982). Nesquehonite<br />

also occurs as evaporative films on magnesium-rich<br />

alkaline wetland waters (Power et al., 2007) and in association<br />

with dypingite and hydromagnesite in lake shore-line<br />

and shallow water carbonate hard grounds (Last et al.,<br />

2010).<br />

Barringtonite has been identified in association with<br />

nesquehonite in precipitates derived from weathered basalts<br />

(Nashar, 1965). Heating nesquehonite at ca 100 °C yields<br />

an unidentified MgCO3 2H2O phase (Ballirano et al.,<br />

2010). Magnesium monohydrate carbonate has also been<br />

synthesised by desiccating nesquehonite (Menzel and<br />

Brückner, 1930). A range <strong>of</strong> [MgCO3 XH2O] phases<br />

(X < 3) have also been synthesised, in which variations in<br />

the structural water content were attributed to varying conditions<br />

<strong>of</strong> synthesis (Zhang et al., 2006). The phases show<br />

mid-infrared bands matching the internal modes <strong>of</strong> the<br />

] anion in nesquehonite, but no X-ray diffraction<br />

[CO 2<br />

3


(XRD) data was reported by the authors. Dihydrate or<br />

monohydrate magnesium carbonates have also been reported<br />

during the solvent-mediated [N ! HM] transition<br />

(e.g., Hopkinson et al., 2008). Complexity in this part <strong>of</strong><br />

the system is compounded by similarities between the thermal<br />

behaviour <strong>of</strong> nesquehonite and hydromagnesite (e.g.,<br />

Beck, 1950; Kazakov et al., 1959; Lanas and Alvarez,<br />

2004), and evidence for the presence <strong>of</strong> bicarbonate in some<br />

synthesised and natural samples <strong>of</strong> nesquehonite. These<br />

attributes have led some investigators to assign a basic magnesium<br />

carbonate formula to nesquehonite: [Mg(OH,<br />

HCO3) 2H2O] (e.g., Wells, 1915; Beck, 1950; Kazakov<br />

et al., 1959; Hales et al., 2008). Hence, either nesquehonite<br />

ranges in composition, or there is a structural isomer <strong>of</strong> the<br />

mineral (Hales et al., 2008).<br />

3. EXPERIMENTS<br />

3.1. Experimental materials and methods<br />

Two experiments were conducted. In both cases, the initial<br />

solution was prepared by adding 3.1 g <strong>of</strong> pulverised<br />

brucite [Mg(OH)2] to 350 ml <strong>of</strong> distilled water at 25 °C<br />

and sparged with pure CO 2 at 1 atm for 19 h until the pH<br />

stabilized at 6.79. At this point, the CO2 flow was stopped<br />

and the temperature increased from 25 to 58 °C (within<br />

10 min). In the first experiment, the system was held at that<br />

temperature for 240 min (static experiment). In the second<br />

experiment, the solution was subject to the same heating<br />

conditions but it was sonicated (agitated experiment). The<br />

rate <strong>of</strong> [H + ] decrease at 58 °C accompanying CO 2 degassing<br />

from pure water previously sparged with CO2 at 25 °C, in<br />

the absence <strong>of</strong> dissolved magnesium, yielded pseudo-first<br />

order rate constants for the decrease in [H + ] <strong>of</strong> k =<br />

0.0225 min 1 when the solution is sonicated and k =<br />

0.0046 min 1 when degassing is exclusively thermally-<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO2–H2O 3<br />

Fig. 1. The hydrated magnesium carbonate minerals in the system CO2–MgO–H2O, adapted from Canterford et al. (1984). Specified values <strong>of</strong><br />

X in synthesised MgCO 3 XH 2O phases include 1.3 and 0.3 H 2O(Zhang et al., 2006). The chemical formula <strong>of</strong> dypingite may show 5 or 6 H 2O<br />

molecules (Xiong and Lord, 2008). The International Mineralogical Association – Commission on New Minerals and Mineral Names (IMA/<br />

CNMNC) status <strong>of</strong> giorgosite is questionable. The status <strong>of</strong> nesquehonite is grandfathered: i.e. the original description preceded the<br />

establishment <strong>of</strong> the CNMNC in 1959, listed formula [MgCO3 3H2O]. The unnamed mineral [Mg5(CO3)4(OH)2 8H2O] is not listed.<br />

driven. The rate equation is expressed as: pHt pH0 =<br />

kt/2.303.<br />

Previous studies show that electric conductance provides<br />

a real-time semi-quantitative method for monitoring carbonate<br />

mineral precipitation events during CO2 degassing<br />

from aqueous solutions (e.g., Zeppenfeld, 2006). In this<br />

study, it is assumed that the electric conductance is largely<br />

due to [Mg 2+ ], [HCO3 ] and [H + ]. Electro-conductivity measurements<br />

were taken using a Jenway 4010 conductivity meter.<br />

Temperature was measured with a Fisher Scientific<br />

platinum sensor (Pt-100O) thermometer (±0.1 °C), while<br />

pH readings were taken with a Mettler Toledo pH meter<br />

(±0.01 pH). The conductivity probe was calibrated with a<br />

standard salt solution supplied by Hanna Instruments,<br />

6.44 parts per thousand KCl at 25 °C, Lot. No B399, calibrated<br />

as 12.88mS, zero was reverse-osmosis water. The pH<br />

probe was calibrated against pH-10 (borate), pH-7 (phosphate)<br />

and pH-4 (phthalate) NIST-traceable buffers purchased<br />

from Fisher Scientific. The temperature correction<br />

for both probes was linear, electric conductivity measurements<br />

were corrected to the 25 °C reference point using a<br />

2.1% per °C coefficient. Readings were acquired from<br />

mid-water depth in both experiments. The precision and<br />

detection limit were ±0.2 mS and ca 0.6 mS respectively.<br />

The magnesium concentration in the solutions was measured<br />

by atomic absorption spectros<strong>copy</strong> (AAS) using a Perkin<br />

Elmer AAnalyst 200 instrument. Small aliquots (2 ml) <strong>of</strong><br />

the experimental solutions were withdrawn from mid-water<br />

depths, filtered through a cellulose acetate membrane<br />

(0.2 lm pore size) and diluted 1:100 with water containing<br />

1 ml <strong>of</strong> spectroscopic grade nitric acid. This solution was further<br />

diluted as appropriate so that the final magnesium concentration<br />

fell within the calibration range <strong>of</strong> (0.2–1.2 ppm)<br />

<strong>of</strong> the instrument. The limit <strong>of</strong> detection was ca 0.02 ppm.<br />

The experiments were repeated several times and halted<br />

at different times for solid phase sample collection and to


verify the reproducibility <strong>of</strong> conductance and pH readings.<br />

In the case <strong>of</strong> the static experiment, precipitation was<br />

marked first by deposition <strong>of</strong> solids at the base <strong>of</strong> the vessel<br />

and then progressive development <strong>of</strong> a surface film <strong>of</strong> acicular<br />

crystals. The two precipitate types were homogenised<br />

during sampling. In the agitated experiment, precipitates remained<br />

suspended in the parent solution. After filtration<br />

through a 0.2 lm filter, powders were air dried at 25 °C<br />

for three days, at constant humidity, then lightly ground<br />

prior to analysis. Samples were labelled with the prefix<br />

[A x] for the agitated experiment and [S x] for the static<br />

experiment, where x = time in minutes at which the sample<br />

was taken after the onset <strong>of</strong> heating. Samples labelled [AS0]<br />

were collected subsequent to CO 2 sparging, and prior to<br />

heating and sonication.<br />

3.2. Solid phase analysis<br />

Fourier-transform mid-infrared (FT-IR), FT-Raman,<br />

and scanning electron microscope analyses (SEM) were<br />

conducted at the <strong>University</strong> <strong>of</strong> <strong>Brighton</strong> (UK). FT-Raman<br />

analyses were undertaken with a 1064 nm Nd-YAG laser<br />

source using a Nicolet Nexus FT-Raman module. Sharp<br />

bands are accurate to 2 cm 1 . FT-IR analyses were performed<br />

using a Nicolet Avatar 320 FT-IR, fitted with a diamond<br />

attenuated total reflectance (ATR) accessory. The<br />

detection limit for mineral phase(s) in mixed assemblages<br />

<strong>of</strong> basic and hydrate magnesium carbonates is 65%, for<br />

FT-Raman and ca 10% for FT-IR. Electron microscopic<br />

analyses were conducted with a Jeol JSM6310 scanning<br />

electron microscope (SEM).<br />

X-ray powder diffraction data (performed at the Natural<br />

History Museum, London, UK) were collected using an<br />

INEL curved position sensitive detector (PSD). This detector<br />

has an output array <strong>of</strong> 4096 digital channels representing<br />

an arc <strong>of</strong> 120°2h and permits the simultaneous<br />

measurement <strong>of</strong> diffracted X-ray intensities at all angles<br />

<strong>of</strong> 2h across 120° with a static beam-sample-detector geometry.<br />

Copper Ka1 radiation was selected from the primary<br />

beam using a germanium 111 single-crystal monochromator.<br />

Horizontal and vertical slits were used to restrict the<br />

beam to 0.14 by 5.0 mm respectively. Powdered samples<br />

were mounted on a single-crystal sapphire substrate, and<br />

measurements made in reflection geometry with the sample<br />

surface (spinning in its own plane) at an angle <strong>of</strong> 2° to the<br />

incident beam. Data collection times were 15 min for each<br />

sample. NIST Silicon powder SRM640 and Silver Behenate<br />

were used as external 2h calibration standards and the 2h<br />

linearization <strong>of</strong> the detector was performed using a leastsquares<br />

cubic spline function. The detection limit for mineral<br />

phase(s) in mixed assemblages <strong>of</strong> basic and hydrate<br />

magnesium carbonates is ca 5% modal abundance.<br />

4. RESULTS<br />

4.1. Electrical conductivity, AAS and pH<br />

Small amounts <strong>of</strong> precipitate formed during sparging <strong>of</strong><br />

CO2 at 25 °C (samples [AS0]). AAS analysis indicated a<br />

corresponding reduction in [Mg 2+ ] from 0.15 to 0.138 M<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

4 L. Hopkinson et al. / Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

during the nineteen hours <strong>of</strong> CO 2 sparging. The onset <strong>of</strong><br />

heating in both experiments produced electrical conductance<br />

data which divides into temporal domains. The first<br />

16 min are defined by a steep decline in electrical conductivity<br />

in both experiments (Fig. 2A), consistent with supersaturation,<br />

solid phase nucleation, and subsequent growth<br />

(e.g., Söhnel and Mullin, 1978; Zeppenfeld, 2006). In the<br />

agitated setting, between ca 16 and 40 min, there is a second<br />

large decrease in electric conductivity, consistent with<br />

a second pulse <strong>of</strong> mineral formation. The [Mg 2+ ] is reduced<br />

to 0.074 M at 25 min, increasing to 0.083 M at 35 min,<br />

thereafter declining to 0.076 M after 120 min (Fig. 2A).<br />

In the static environment, within the same time frame, a<br />

modest decrease in conductivity and [Mg 2+ ] is observed<br />

(Fig. 2A), consistent with slow growth <strong>of</strong> crystals in solution<br />

as supersaturation was depleted (e.g., Söhnel and Mullin,<br />

1978).<br />

Within the first ten minutes <strong>of</strong> the experiments, the pH<br />

increases at a faster rate in the agitated environment than<br />

the static environment (Fig. 2B), consistent with accelerated<br />

CO 2 loss in the former. Between 15 and 20 min, the pH in<br />

the agitated environment undergoes an abrupt reduction<br />

from 7.93 to 7.86, reflecting CO2 production resulting from<br />

carbonate mineral formation. Between 20 and 120 min, the<br />

pH in the agitated setting shows little variation, measured<br />

at pH = 7.98 at the 120-min mark. In contrast, the pH in<br />

the static environment undergoes a protracted gradual increase,<br />

achieving parity with the agitated environment readings<br />

after ca 35 min and pH = 8.09 120 min into the<br />

experiment (Fig. 2B). Monitoring <strong>of</strong> the static and agitated<br />

experiments continued intermittently for up to 240 min. No<br />

appreciable changes in pH or electric conductance, relative<br />

to the 120-min readings occurred in this time frame.<br />

4.2. X-ray diffraction<br />

[AS 0] precipitates are identified as hydromagnesite,<br />

[Mg5(CO3)4(OH)2 4H2O] (Fig. 3A). Samples [A20], [A50],<br />

and [A80] are dominated by nesquehonite with subordinate<br />

hydromagnesite and traces <strong>of</strong> a dypingite-like mineral<br />

phase(s) for which the closest match is [Mg 5(CO 3) 4(OH) 2<br />

5H2O] (Fig. 3B and C). Note that discrimination between<br />

[MgCO3 3H2O] and [Mg(HCO3,OH) 2H2O] is not possible<br />

by XRD. Sample [A 120] is dominated by hydromagnesite,<br />

includes a dypingite-like phase(s), but is devoid <strong>of</strong> nesquehonite.<br />

The diffraction pattern shows two low-angle peaks<br />

not present in dypingite (Fig. 3C). Only the reference diffraction<br />

pattern <strong>of</strong> the unnamed dypingite-like phase,<br />

[Mg5(CO3)4(OH)2 8H2O] (Suzuki and Ito, 1973) contains<br />

these peaks, <strong>of</strong> which, the peak at 2.7° (33 A ˚ d-spacing) is<br />

the main distinguishing feature relative to the 5H2O version<br />

<strong>of</strong> dypingite and hydromagnesite (Fig. 3). Given that<br />

Suzuki and Ito (1973) present a water determination, the<br />

experimental data is in keeping with the unnamed mineral.<br />

The multi-minerallic make up <strong>of</strong> the precipitates, combined<br />

with extensive overlap <strong>of</strong> diffraction patterns between<br />

dypingite and the unnamed mineral dictates that the simultaneous<br />

presence <strong>of</strong> dypingite is not precluded.<br />

Sample [A240] gives a diffraction pattern consistent with<br />

an additional dypingite-like mineral phase with subordinate


hydromagnesite. The dypingite-like structure appears to<br />

have a smaller cell volume than the unnamed mineral and<br />

the 5H2O version <strong>of</strong> dypingite, as all peaks are shifted to<br />

smaller d-spacings (Fig. 3E). Hereafter, excluding hydromagnesite,<br />

[Mg5(CO3)4(OH)2 XH2O] minerals where<br />

X = 11, 8, 6, 5 or less are collectively referred to as dypingite-type<br />

mineral phases. The full-width at half-maximum<br />

values (FWHM) <strong>of</strong> the dypingite-type minerals are wider<br />

than hydromagnesite, suggesting greater long-range disorder<br />

in the dypingite-type minerals. Hydromagnesite shows<br />

similar FWHM throughout, even though the peaks have<br />

different relative intensities, as monitored by the peak at<br />

9.6 degrees. The static experiment precipitates show<br />

nesquehonite-rich hydromagnesite bearing assemblages.<br />

There is no evidence <strong>of</strong> a shift in the abundance <strong>of</strong> nesquehonite<br />

relative to hydromagnesite with time, or the appearance<br />

<strong>of</strong> dypingite-type phases.<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO2–H2O 5<br />

Fig. 2. Experimental results for: (A) electrical conductance measurements (agitated environment symbol: , static environment symbol:<br />

). AAS analyses <strong>of</strong> [Mg 2+ ]tot/(M) concentrations (static environment symbol: , agitated environment symbol: ); (B) pH measurements,<br />

(agitated environment symbol: , static environment symbol: ). Agitated environment solid phase samples are: [AS0], [A20], [A50], [A80],<br />

[A 120] and [A 240]. Static environment samples are: [S 60], [S 120], [S 180], and [S 240].<br />

4.3. FT-Raman<br />

All FT-Raman spectra contained significant noise, particularly<br />

those acquired from the agitated experiment. This<br />

relates to the acknowledged difficulty in analysing disordered<br />

magnesium hydrate and hydroxyl carbonates (e.g.,<br />

White, 1974; Frost et al., 2009), compounded by the rapid<br />

synthesis under our experimental conditions. The [AS0]<br />

spectra (Fig. 4A) show a high intensity band at 1117 cm 1 ,<br />

consistent with the v1 internal mode <strong>of</strong> [CO 2<br />

3 ] for hydromagnesite<br />

(e.g., Edwards et al., 2005) and a range <strong>of</strong> low<br />

intensity bands assigned to hydromagnesite. In the 2500–<br />

3800 cm 1 (H2O–OH) region, [AS0] shows an asymmetric<br />

band at ca 3660 cm 1 . The spectra <strong>of</strong> [AS0] samples collected<br />

after 5, 8, and 19 h <strong>of</strong> CO2 sparging are identical, suggesting<br />

that hydromagnesite precipitated directly from the<br />

parent solution.


The [A20] and [A50] spectra are very similar, showing a<br />

high intensity band at 1098 cm 1 , with a shoulder at<br />

1117 cm 1 , plus a variety <strong>of</strong> low intensity bands assigned<br />

to nesquehonite and or hydromagnesite (e.g., Fig. 4B).<br />

Hence, results <strong>of</strong> the FT-Raman analysis support XRD evidence<br />

for a transition from hydromagnesite to nesquehonite<br />

formation. The spectra show broad bands centred at ca<br />

1480 and 1780 cm 1 . Bands at similar frequencies have<br />

been assigned to the anti-symmetric stretching <strong>of</strong> [CO 2<br />

3 ]<br />

and the water bending mode for nesquehonite (Hales<br />

et al., 2008). In the H2O–OH region, the asymmetric feature<br />

between ca 3000–3500 cm 1 contains poorly resolved overlapping<br />

bands at ca 3440, 3310 and 3156 cm 1 , the disposition<br />

<strong>of</strong> which are broadly compatible with synthetic<br />

[Mg(HCO3,OH) 2H2O] (e.g., Hales et al., 2008). A subsample<br />

<strong>of</strong> [A20] was analysed periodically over five days<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

6 L. Hopkinson et al. / Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

Fig. 3. X-ray diffraction pattern <strong>of</strong> samples: (A) [AS 0]; (B) [A 20]; (C) [A 80]; (D) [A 120]; (E) [A 240]. The star marked on the stick diagram <strong>of</strong><br />

[Mg5(CO3)4(OH)2 8H2O], is the main peak enabling discrimination <strong>of</strong> the mineral phase from dypingite and hydromagnesite.<br />

<strong>of</strong> drying. The 1098/1117 cm 1 intensity ratio was uniform,<br />

suggesting that the loss <strong>of</strong> nesquehonite at the expense <strong>of</strong><br />

[Mg5(CO3) 4(OH) 2 XH2O] phases was solely solventmediated.<br />

The [A80] spectrum shows an elevated (1117/1098 cm 1 )<br />

intensity ratio relative to [A20] and [A50](Fig. 4C), consistent<br />

with the presence <strong>of</strong> hydromagnesite generated at 25 °C,<br />

combined with nesquehonite dissolution, and coeval formation<br />

<strong>of</strong> dypingite-type phases. The [A80] and [A120] spectra<br />

show broad low intensity asymmetric scattering in the ca<br />

1000–1090 cm 1 region which is multi-component in nature<br />

(Fig. 4D). The lower frequency region coincides with Mg–<br />

OH deformation vibrations in dypingite (Frost et al.,<br />

2009). Low intensity scattering in the 1070 cm 1 region is<br />

consistent with the [CO 2<br />

3 ] symmetric stretching mode<br />

(Frost et al., 2008). In keeping with XRD analysis, samples


[A 120] and [A 240] are devoid <strong>of</strong> nesquehonite. The spectra<br />

also show a low intensity sharp band at 1092 cm 1<br />

(Fig. 4D and E). A similar feature has been described from<br />

natural dypingite (Frost et al., 2009). The width <strong>of</strong> broad<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO2–H2O 7<br />

Fig. 4. FT-Raman spectra <strong>of</strong> samples: (A) [AS0]; (B) [A20]; (C) [A80]; (D) [A120]; (E) [A240]; (F) [S240]. [N] denotes a nesquehonite band<br />

assignment; [HM] a hydromagnesite band assignment; [Dy-t] a dypingite-type mineral band assignment. The black diamond present in<br />

Mg 5(CO 3) 4(OH) 2 XH 2O-rich assemblages marks the area <strong>of</strong> broad band absorption in the ca 1000–1090 cm 1 region. See text for details.<br />

low intensity scattering on the shoulder <strong>of</strong> the high intensity<br />

1120 cm 1 band is narrower in [A240] than in [A120].<br />

Spectra <strong>of</strong> precipitates formed under static conditions<br />

complement XRD data, showing a variety <strong>of</strong> bands


assigned to nesquehonite and/or hydromagnesite (e.g.,<br />

Fig. 4F). No increase in the (1117/1099 cm 1 ) intensity ratio<br />

with increasing heating time is evident, suggesting no<br />

change in the abundance <strong>of</strong> nesquehonite relative to hydromagnesite<br />

with increasing time. Nevertheless, the intensity<br />

ratio is elevated relative to the [A20], [A50] and [A80] spectra,<br />

consistent with greater nesquehonite generation in the agitated<br />

setting. In the H 2O–OH region, overlapping bands<br />

at ca 3156, 3312 and 3437 cm 1 are consistent with water<br />

stretching vibrations for synthetic [Mg(HCO3,OH) 2H2O]<br />

(Hales et al., 2008). There is no evidence for broad band<br />

scattering in the ca 950–1090 cm 1 region, suggesting that<br />

this spectral attribute is singular to the dypingite-type<br />

phases generated in the agitated setting.<br />

4.4. FT-mid infrared<br />

The [AS0] spectrum (Fig. 5A) shows a series <strong>of</strong> bands assigned<br />

to the [CO 2<br />

3 ] anion in hydromagnesite (e.g., Lanas<br />

and Alvarez, 2004): 1117 cm 1 (symmetrical stretching<br />

vibration); 793, 852 and 885 cm 1 (bending vibrations),<br />

1420 and 1477 cm 1 (anti-symmetric stretching vibrations).<br />

Weak absorption at ca 1660 cm 1 (H2O bending vibration)<br />

is also compatible with hydromagnesite (e.g., White, 1974).<br />

The spectrum also shows a broad concave spectral feature<br />

at ca 1000 cm 1 , which occurs in some hydromagnesite<br />

and dypingite spectra (e.g., Raade, 1970; White, 1971). This<br />

feature has previously been assigned to deformation modes<br />

<strong>of</strong> Mg–OH units (Frost et al., 2008) and solid phase incorporation<br />

<strong>of</strong> bicarbonate (e.g., Zhang et al., 2006). In the<br />

H2O–OH region, [AS0] is comparable with published spectra<br />

<strong>of</strong> hydromagnesite (e.g., White, 1971), with absorption<br />

at ca 3038 and 3444 cm 1 (H2O stretching vibrations) and<br />

super-imposed sharp ([OH ] stretching vibration) bands<br />

at 3510 and 3650 cm 1 .<br />

Samples [A20] and [A50] show bands at 1098 and<br />

852 cm 1 (e.g., Fig. 5B), consistent with symmetric stretching,<br />

and bending <strong>of</strong> the [CO 2<br />

3 ] anion for nesquehonite (e.g.,<br />

White, 1974; Lanas and Alvarez, 2004; Zhang et al., 2006;<br />

Ferrini et al., 2009). Weak absorption at 888 cm 1 is also<br />

evident, consistent with XRD and FT-Raman evidence<br />

for hydromagnesite, produced at 25 °C in subordinate concentrations<br />

to nesquehonite. The spectra also show bands<br />

at ca 1420, 1471, and 1519 cm 1 in keeping with the antisymmetric<br />

stretching mode <strong>of</strong> [CO 2<br />

3 ] for nesquehonite,<br />

superimposed on anti-symmetric stretching vibrations <strong>of</strong><br />

hydromagnesite. Resolution <strong>of</strong> the individual anti-symmetric<br />

stretching modes <strong>of</strong> nesquehonite also varies in separate<br />

studies (e.g., Coleyshaw et al., 2003; Kloprogge et al., 2003;<br />

Zhang et al., 2006). Moderate absorption at ca 1648 cm 1 is<br />

assigned to the superimposed OH bending mode <strong>of</strong> H2O for<br />

nesquehonite and hydromagnesite (White, 1971).<br />

In the H2O–OH region, [A20] and [A50] show a sharp<br />

band at 3555 cm 1 and over-lapping bands at ca 3430,<br />

3246 and 3122 cm 1 (Fig 5b). The spectra are dissimilar<br />

to [MgCO3 XH2O] phases (Zhang et al., 2006) in which<br />

no sharp band absorption occurs at 3555 cm 1 . The four<br />

bands do, however, broadly coincide with Raman active<br />

bands <strong>of</strong> synthesised [Mg(HCO3,OH) 2H2O], measured at<br />

3124, 3295, 3423 and 3550 cm 1 <strong>of</strong> which the first three<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

8 L. Hopkinson et al. / Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

are assigned to water stretching vibrations and the latter<br />

to the stretching mode <strong>of</strong> OH units (Hales et al., 2008).<br />

The samples also show broad low intensity band(s) in the<br />

ca 2500 cm 1 region which partially overlaps with the characteristic<br />

frequency range (2200–2500 cm 1 ) for the bicarbonate<br />

ion (White, 1971).<br />

Samples [A80] and [A120] contain bands which coincide<br />

with the internal modes <strong>of</strong> the [CO 2<br />

3 ] anion for hydromag-<br />

nesite and dypingite (Fig. 5C and D). Also evident is an increase<br />

in 888 cm 1 absorption intensity relative to [A20] and<br />

[A 50], consistent with a reduction in the ratio <strong>of</strong> nesquehonite<br />

to hydromagnesite and dypingite-type phases with<br />

increasing heating time. Given that the dominant (or sole)<br />

dypingite-type phase identified by XRD in [A 120] is<br />

[Mg5(CO3)4(OH)2 8H2O], it seems that the unnamed mineral,<br />

like dypingite, shows short-range order akin to hydro-<br />

magnesite, with respect to the internal modes <strong>of</strong> the [CO 2<br />

3 ]<br />

anion. Sample [A 240], which shows XRD evidence for a second<br />

dypingite-type phase, similarly possesses short-range<br />

order <strong>of</strong> the carbonate anion akin to hydromagnesite.<br />

In the H 2O–OH region, samples [A 80], [A 120] and [A 240]<br />

show the progressive development <strong>of</strong> broad absorption at<br />

ca 2950 and 3430 cm 1 , assigned to water stretching bands,<br />

with super-imposed sharp bands at 3510 and 3650 cm 1 ,<br />

comparable in frequency with hydromagnesite and some<br />

dypingite spectra, in which the bands are assigned to<br />

stretching vibrations <strong>of</strong> OH units (e.g., White, 1974; Frost<br />

et al., 2008). Samples [A80], [A120] and [A240] show variably<br />

resolved broad, low-intensity absorption centered at ca<br />

1000 cm 1 (Fig. 5C–E). The feature coincides with two<br />

bands (at 948 and 1012 cm 1 ) identified in dypingite and assigned<br />

to deformation modes <strong>of</strong> Mg–OH units (Frost et al.,<br />

2008). The feature is also present in some published spectra<br />

<strong>of</strong> hydromagnesite (e.g., White, 1971).<br />

The spectra <strong>of</strong> precipitates formed under static conditions<br />

show a range <strong>of</strong> bands assigned to nesquehonite<br />

and/or hydromagnesite (e.g., Fig. 5F), with the latter mineral<br />

contributing more strongly to the spectra, than in samples<br />

[A20], [A50], [A80] recovered from the agitated<br />

experiment. In the H 2O–OH region <strong>of</strong> samples formed under<br />

static conditions, spectra are similar to nesquehoniterich<br />

spectra <strong>of</strong> samples generated in the agitated environment.<br />

Weak absorption at 3655 cm 1 is assigned to OH<br />

units in hydromagnesite.<br />

4.5. SEM<br />

Samples [AS0] consist <strong>of</strong> platy (ca 1–2 lm length) hydromagnesite<br />

crystals, organised in agglomerates (Fig. 6A).<br />

Samples [A 20] and [A 50] are dominated by columnar<br />

nesquehonite crystals, ca 5–30 lm in the longest dimension<br />

(Fig. 6B). Also present are subordinate quantities <strong>of</strong> hydromagnesite<br />

agglomerates. Sample [A 80] consists <strong>of</strong> nesquehonite<br />

rods, showing etch pits and overgrowths <strong>of</strong> platy<br />

basic carbonates, producing a ‘house <strong>of</strong> cards’ texture<br />

(Fig. 6C). The texture is believed to be related to a dissolution–recrystallization<br />

self-assembly growth mechanism, in<br />

which unstable, dissolving nesquehonite micro-rods function<br />

as templates for hydromagnesite, which in turn act as<br />

nucleation points for further hydromagnesite platelets


(Hao and Du, 2009). Development <strong>of</strong> the texture hinges on<br />

the balance between the dissolution rate <strong>of</strong> nesquehonite<br />

and the precipitation rate <strong>of</strong> hydromagnesite, the latter process<br />

occurring at a slower rate than nesquehonite dissolution<br />

(Hao and Du, 2009). Given that dypingite-type<br />

phases appear to grow at the expense <strong>of</strong> nesquehonite<br />

and that [A120] shows pronounced house <strong>of</strong> cards textural<br />

development, in the absence <strong>of</strong> nesquehonite and the pres-<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO2–H2O 9<br />

Fig. 5. Mid-infrared spectra <strong>of</strong> experimental precipitates, for samples: (A) [AS0]; (B) [A20]; (C) [A80]; (D) [A120]; (E) [A240]; (F) [S240]. Between<br />

1900 and 2400 cm 1 the spectra contain ATR related diamond spectra (e.g., Grice et al., 1991). [N] denotes a nesquehonite band assignment;<br />

[HM] a hydromagnesite band assignment; [Dy-t] a dypingite-type band assignment. The black diamond present in Mg 5(CO 3) 4(OH) 2 XH 2Orich<br />

assemblages marks the area <strong>of</strong> broad band absorption in the ca 1000–1090 cm 1 region. See text for details.<br />

ence <strong>of</strong> dypingite-type phase(s) (Fig. 6D), it follows that the<br />

growth mechanism may be applicable to a range <strong>of</strong> dypingite-type<br />

phases. Static environment samples are dominated<br />

by nesquehonite rods, which range from ca 50–200 lm in<br />

the longest dimension, in association with small quantities<br />

<strong>of</strong> platy agglomerates <strong>of</strong> carbonates texturally akin to<br />

[AS0]. The samples are devoid <strong>of</strong> evidence for overgrowth<br />

<strong>of</strong> nesquehonite in the form <strong>of</strong> platy basic carbonates, or


evidence <strong>of</strong> partial nesquehonite dissolution, such as etch<br />

pits (Fig. 6E and F).<br />

5. DISCUSSION<br />

Fig. 7 shows the solution saturation indices for mineral<br />

phases relevant to this study, calculated using Geochemists<br />

Workbench Ò and data from the agitated experiment. The<br />

model indicates that the solution was close to saturation<br />

with respect to brucite and supersaturated with respect to<br />

magnesite throughout the 120 min <strong>of</strong> reaction at 58 °C.<br />

At room temperature, the solution is assumed to be in equilibrium<br />

with hydromagnesite and is undersaturated with<br />

nesquehonite. Subsequently, thermally-accelerated CO 2<br />

degassing rapidly promoted the precipitation <strong>of</strong> nesquehonite<br />

[Mg(HCO3,OH) 2H2O]. At this juncture, the parent<br />

solution was simultaneously slightly supersaturated with respect<br />

to nesquehonite and strongly supersaturated with respect<br />

to hydromagnesite. Nevertheless, our experimental<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

10 L. Hopkinson et al. / Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

Fig. 6. Back-scattered images <strong>of</strong> experimental precipitates. (A) Agglomerate <strong>of</strong> hydromagnesite crystals from sample [AS0]. (B) Acicular<br />

nesquehonite crystals with subordinate hydromagnesite agglomerates Sample [A 20]. (C) Overgrowths <strong>of</strong> platy carbonates on nesquehonite<br />

rod-shaped crystals [A 80]. XRD indicates the presence <strong>of</strong> nesquehonite, dypingite-type phase(s) and hydromagnesite in the sample. (D) Sample<br />

[A120] displaying a well-developed house <strong>of</strong> cards texture. Sample [A240] is texturally indistinguishable from [A120]. (E) Sample [S120] shows<br />

nesquehonite crystals in association with agglomerates <strong>of</strong> platy crystals texturally comparable to sample [AS0]. (F) Sample [S240] is texturally<br />

comparable to [S 120], with no evidence for development <strong>of</strong> platelets on nesquehonite surfaces.<br />

results indicate that nesquehonite was kinetically favoured.<br />

This is presumably because metastable reaction products<br />

with simpler structures form more rapidly than the more<br />

complicated although thermodynamically more stable<br />

phase (e.g., Goldsmith, 1953; Morse and Casey, 1988) even<br />

though the system was also seeded with hydromagnesite.<br />

Hence, growth <strong>of</strong> hydromagnesite directly from the parent<br />

solution most likely continued, albeit at a greatly reduced<br />

rate relative to nesquehonite. Results show a greater yield<br />

<strong>of</strong> nesquehonite in the agitated setting and the finer size<br />

range <strong>of</strong> nesquehonite precipitates relative to precipitates<br />

obtained from the static environment. The short-lived negative<br />

excursion in the pH in the agitated experiment likely<br />

results from the enhanced production <strong>of</strong> CO 2 upon precipitation<br />

<strong>of</strong> the carbonate mineral.<br />

The multi-minerallic make up <strong>of</strong> the precipitates dictates<br />

that no unambiguous conclusion concerning the presence <strong>of</strong><br />

[MgCO 3 3H 2O] can be drawn. Nevertheless, available<br />

evidence suggests that, where nesquehonite occurs in


abundance [Mg(HCO3,OH) 2H2O] is dominant, possibly to<br />

the exclusion <strong>of</strong> the tri-hydrate. It seems unlikely that the<br />

temperature <strong>of</strong> synthesis plays a pivotal role in determining<br />

whether basic or hydrate nesquehonite forms, as both variants<br />

have been synthesised at low temperatures, and both<br />

occur in natural near-surface ambient temperature settings<br />

(e.g., Kazakov et al., 1959; Coleyshaw et al., 2003; Hales<br />

et al., 2008). Accordingly, it is plausible that the pH influences<br />

the extent to which HCO 3 is incorporated into<br />

nesquehonite or which isomer <strong>of</strong> nesquehonite is produced.<br />

In this respect, it is interesting to note that the synthesis <strong>of</strong><br />

nesquehonite [MgCO3 XH2O] phases by Zhang et al. (2006)<br />

occurred at pH values <strong>of</strong> 8.5–12.5. In contrast, in the<br />

experiment documented here, nesquehonite synthesis was<br />

achieved at pH < 8, yet the temperatures <strong>of</strong> synthesis and<br />

timeframes for mineral formation in the two studies are<br />

comparable.<br />

The solubility <strong>of</strong> dypingite-type phases is not known<br />

with any degree <strong>of</strong> certainty. Nevertheless, XRD results<br />

indicate that the system was supersaturated with dypingite-type<br />

phase(s) 20 min after the beginning <strong>of</strong> the heating<br />

stage, coincident with the onset <strong>of</strong> very low levels <strong>of</strong> undersaturation<br />

in nesquehonite (Fig. 7). The progressive emergence<br />

<strong>of</strong> dypingite-type phases is associated with particles<br />

heterogeneously nucleating on decomposing nesquehonite,<br />

giving rise to house <strong>of</strong> cards textures. Heterogeneous nucleation<br />

may be responsible for Ostwald step rule behaviour,<br />

as the next most stable phase is <strong>of</strong>ten more structurally similar<br />

to the precursor phase than a thermodynamically more<br />

stable phase (Morse and Casey, 1988).<br />

Nesquehonite dissolution and dypingite-type mineral<br />

formation are more extensive in the agitated environment,<br />

relative to the static environment. Strong hydrodynamic<br />

shear forces generated by sonication can increase the rate<br />

<strong>of</strong> dissolution <strong>of</strong> suspended solids by de-agglomeration,<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO2–H2O 11<br />

Fig. 7. Agitated experiment saturation indices. Note, there are no available thermodynamic data for dypingite-type phases. See text for<br />

details.<br />

simultaneously accelerating the formation <strong>of</strong> viable nuclei<br />

to increase the rate <strong>of</strong> crystallization <strong>of</strong> carbonate mineral<br />

phases (e.g., Kim et al., 2011). No protohydromagnesite<br />

was identified in this study, either because it was rapidly<br />

superseded by [Mg5(CO3)4(OH)2 XH2O] phases, or because<br />

its formation was prohibited by the nature <strong>of</strong> the nesquehonite<br />

precursor. This may be due to conditions <strong>of</strong> synthesis<br />

or incongruent water loss prior to complete dissolution.<br />

Samples [AS0] (hydromagnesite), the static environment<br />

powders, and samples [A80], [A120] and [A240] (dypingitetype<br />

and hydromagnesite bearing) all show pronounced<br />

broad band infrared absorption in the ca 1000 cm 1 region,<br />

assigned in large measure to Mg(OH) deformation modes.<br />

The hydromagnesite spectra [AS0] and static environment<br />

spectra are devoid <strong>of</strong> the corresponding Raman active<br />

band(s), although the bands are clearly resolved in the<br />

FT-Raman spectra <strong>of</strong> the three precipitates formed in the<br />

agitated environment. These attributes are consistent with<br />

reduced symmetry and therefore greater disorder in the<br />

dypingite-type phases relative to hydromagnesite and, thus,<br />

the greater ease <strong>of</strong> crystallization <strong>of</strong> these phases relative to<br />

hydromagnesite. The greater numbers <strong>of</strong> waters <strong>of</strong> crystallization<br />

<strong>of</strong> [Mg5(CO3) 4(OH) 2 8H2O] relative to hydromagnesite<br />

evidently imparts distinct unit cell parameters,<br />

simultaneously affecting Mg(OH) deformation modes, yet<br />

retaining essentially uniform short-range order <strong>of</strong> the<br />

[CO 2<br />

3 ] anion with respect to hydromagnesite. The reduction<br />

in FT-Raman intensity <strong>of</strong> scattering in the ca<br />

1000 cm 1 region in [A240] relative to [A80] and [A120] is<br />

in keeping with decreasing disorder <strong>of</strong> dypingite-type<br />

phases with increasing heating (reaction) time, whereas<br />

the smaller d-spacing <strong>of</strong> [A240] relative to [A120] is attributed<br />

to cell shrinkage with decreasing waters <strong>of</strong> crystallization.<br />

Experimental results indicate that, with increasing reaction<br />

time, small amounts <strong>of</strong> dypingite in association with


nesquehonite in [A 20] are followed by significant quantities<br />

<strong>of</strong> [Mg5(CO3)4(OH)2 8H2O] and, in turn, progressively<br />

superseded by dypingite-type minerals, with waters <strong>of</strong> crystallization<br />

<strong>of</strong> five or less, accompanied by the loss <strong>of</strong> resolvable<br />

nesquehonite. These attributes are consistent with the<br />

solvent-mediated [N ! HM] transition proceeding through<br />

the generation <strong>of</strong> multiple dypingite-type intermediates, in<br />

which the rates <strong>of</strong> formation and transformation <strong>of</strong> metastable<br />

intermediate to metastable intermediate changes with<br />

time. The overall decrease in solid phase disorder and water<br />

content with increasing time is in keeping with the generation<br />

<strong>of</strong> dypingite-type phases providing a mechanism which<br />

serves to minimize entropy production. Thereby, the<br />

[N ! HM] transition obeys the Ostwald (1897) step rule<br />

for successive reactions (e.g., van Santen, 1984).<br />

6. CONCLUSIONS<br />

The solvent-mediated [N ! HM] transition is shown to<br />

be mixed diffusion and reaction controlled, with dypingitetype<br />

phases initially developing from the dissolution products<br />

<strong>of</strong> nesquehonite. The temporal existence <strong>of</strong> different<br />

dypingite-type phase dominated assemblages is consistent<br />

with the transition involving a suite <strong>of</strong> reactions for which<br />

the sequence <strong>of</strong> reaction intermediates is controlled by the<br />

reaction rates, partly accounting for the variety <strong>of</strong> hydromagnesite-like<br />

[Mg 5(CO 3) 4(OH) 2 XH 2O] phases, reported<br />

in separate studies <strong>of</strong> the [N ! HM] transition (e.g., Davies<br />

and Bubela, 1973). The system was highly supersaturated<br />

and seeded with hydromagnesite, throughout the period<br />

<strong>of</strong> nesquehonite formation and transformation. Hence, it<br />

is likely that hydromagnesite continued to form directly<br />

from the parent solution and indirectly via the [N ! HM]<br />

transition. Accepting this, it follows that parallel reaction<br />

pathways existed, with the occurrence <strong>of</strong> nesquehonite<br />

and dypingite-type phases controlled by the rates <strong>of</strong> the direct<br />

and indirect pathways to hydromagnesite.<br />

Given that dypingite with five or less waters <strong>of</strong> crystallization<br />

supersedes the metastable intermediate with eight<br />

waters <strong>of</strong> crystallization, it is possible that dypingite-type<br />

phases lie progressively further from the free energy <strong>of</strong> hydromagnesite<br />

with increasing compositional approach to<br />

nesquehonite. Nevertheless, nesquehonite varies in composition,<br />

be it imparted by conditions <strong>of</strong> synthesis (e.g., Zhang<br />

et al., 2006) or variable desiccation (e.g., Menzel and<br />

Brückner, 1930). Further, the degree <strong>of</strong> congruency <strong>of</strong><br />

nesquehonite dissolution is uncertain. Therefore, the overall<br />

[N ! HM] transition may not be a simple series <strong>of</strong> reactions,<br />

in which all possible dypingite-type intermediates<br />

are generated.<br />

ACKNOWLEDGEMENTS<br />

Carbon Connections are thanked for financial support, Grant<br />

0038. Dr. A. Mucci is thanked for his editorial handling <strong>of</strong> this<br />

work, which has led to its significant improvement. Two anonymous<br />

reviewers and Dr. M. Prieto are likewise thanked for their<br />

significant positive contributions to this study. Dr. Martin Smith<br />

is also gratefully thanked for his involvement.<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

12 L. Hopkinson et al. / Geochimica et Cosmochimica Acta 76 (2012) 1–13<br />

REFERENCES<br />

Ballirano P., De Vito C., Ferrini V. and Mignardi S. (2010) The<br />

thermal behaviour and structural stability <strong>of</strong> nesquehonite,<br />

MgCO3 3H2O, evaluated by in situ laboratory parallel-beam Xray<br />

powder diffraction: new constraints on CO2 sequestration<br />

within minerals. J. Hazard. Mater. 178, 522–528.<br />

Beck C. W. (1950) Differential thermal analysis curves <strong>of</strong> carbonate<br />

minerals. Am. Mineral. 12, 985–1013.<br />

Botha A. and Strydom C. A. (2001) Preparation <strong>of</strong> magnesium<br />

hydroxy carbonate from magnesium hydroxide. Hydrometallurgy<br />

62, 175–183.<br />

Canterford J. H., Tsambourakis G. and Lambert B. (1984) Some<br />

observations on the properties <strong>of</strong> dypingite Mg5(CO3)4(OH)2<br />

5H 2O, and related minerals. Mineral. Mag. 48, 437–442.<br />

Cheng W. and Li Z. (2010) Nucleation kinetics <strong>of</strong> nesquehonite<br />

(MgCO3 3H2O) in the MgCl2–Na2CO3 system. J. Cryst. Growth<br />

312, 1563–1571.<br />

Coleyshaw E. E., Crump G. and Griffith W. P. (2003) Vibrational<br />

spectra <strong>of</strong> the hydrated carbonate minerals ikaite, monohydrocalcite,<br />

lansfordite and nesquehonite. Spectrochim. Acta A59,<br />

2231–2239.<br />

Davies P. J. and Bubela B. (1973) The transformation <strong>of</strong><br />

nesquehonite into hydromagnesite. Chem. Geol. 12, 289–300.<br />

Davies P. J., Bubela B. and Ferguson J. (1977) Simulation <strong>of</strong><br />

carbonate diagenetic processes: formation <strong>of</strong> dolomite, huntite<br />

and monohydrocalcite by the reactions between nesquehonite<br />

and brine. Chem. Geol. 19, 187–214.<br />

Edwards H. G. M., Jorge Villar S. E., Jehlicka J. and Munshi T.<br />

(2005) FT-Raman spectroscopic study <strong>of</strong> calcium-rich and<br />

magnesium-rich carbonate minerals. Spectrochim. Acta 61,<br />

2273–2280.<br />

Ferrini V., De Vito C. and Mignardi S. (2009) Synthesis <strong>of</strong><br />

nesquehonite by reaction <strong>of</strong> gaseous CO 2 with Mg chloride<br />

solution: its potential role in the sequestration <strong>of</strong> carbon<br />

dioxide. J. Hazard. Mater. 168, 832–837.<br />

Friedel B. (1975) Synthetischer giorgiosit. Neues Jahrb. Mineral.<br />

Monatsh. 26, 196–208.<br />

Frost R. L., Bahfenne S., Graham J. and Reddy B. J. (2008) The<br />

structure <strong>of</strong> selected magnesium carbonate minerals – a near<br />

infrared and mid-infrared spectroscopic study. Polyhedron 27,<br />

2069–2076.<br />

Frost R. L., Bahfenne S. and Graham J. (2009) Raman spectroscopic<br />

study <strong>of</strong> the magnesium-carbonate minerals – artinite<br />

and dypingite. J. Raman Spectrosc. 39, 855–860.<br />

Goldsmith J. R. (1953) A “simplexity principle” and its relation to<br />

“ease” <strong>of</strong> crystallization. J. Geol. 61, 439–451.<br />

Grice J. D., Nickel E. H. and Gault R. A. (1991) Ashburtonite, a<br />

new bicarbonate-silicate mineral from Ashburton Downs,<br />

Western Australia: description and structure determination.<br />

Am. Mineral. 76, 1701–1707.<br />

Hales M. C., Frost R. L. and Martens W. N. (2008) Thermo-<br />

Raman spectros<strong>copy</strong> <strong>of</strong> synthetic nesquehonite–implications<br />

for the geosequestration <strong>of</strong> greenhouse gases. J. Raman<br />

Spectrosc. 39, 1141–1149.<br />

Hänchen M., Prigiobbe V., Baciocchi R. and Mazzotti M. (2008)<br />

Precipitation in the Mg-carbonate system – effects <strong>of</strong> temperature<br />

and CO2 pressure. Chem. Eng. Sci. 63, 1012–1028.<br />

Hao Z. and Du F. (2009) Synthesis <strong>of</strong> basic magnesium carbonate<br />

microrods with a ‘house <strong>of</strong> cards’ surface structure using rodlike<br />

particle template. J. Phys. Chem. Solids 70, 401–404.<br />

Hill R. J., Canterford J. H. and Moyle F. J. (1982) New data for<br />

landsfordite. Mineral. Mag. 46, 453–457.<br />

Hopkinson L., Rutt K. and Cressey G. (2008) The nesquehonite<br />

to hydromagnesite transition in the system


MgO CaO H 2O CO 2: an experimental spectroscopic study.<br />

J. Geol. 116, 387–400.<br />

Kazakov A. V., Tikhomirova M. M. and Plotnikova V. I. (1959)<br />

The system <strong>of</strong> carbonate equilibria. Int. Geol. Rev. 1(10), 1–39.<br />

Kim J.-M., Chang S.-M., Kim K.-S., Chung M.-K. and Kim W.-S.<br />

(2011) Acoustic influence on aggregation and agglomeration <strong>of</strong><br />

crystals in reaction crystallization <strong>of</strong> cerium carbonate. Colloids<br />

Surf. A 375, 193–199.<br />

Kloprogge J. T., Martens W. N., Duong L. V. and Webb G. E.<br />

(2003) Low temperature synthesis and characterisation <strong>of</strong><br />

nesquehonite. J. Mater. Sci. Lett. 22(11), 825–829.<br />

Königsberger E., Königsberger L.-C. and Gamsjäger H. (1999)<br />

Low temperature thermodynamic model for the system:<br />

Na2CO3 MgCO3 CaCO3–H2O. Geochim. Cosmochim. Acta<br />

63, 3105–3119.<br />

Lanas J. and Alvarez J. I. (2004) Dolomitic lime: thermal<br />

decomposition <strong>of</strong> nesquehonite. Thermochim. Acta 421, 123–<br />

132.<br />

Langmuir D. (1965) Stability <strong>of</strong> carbonates in the system<br />

MgO CO 2–H 2O. J. Geol. 73, 730–754.<br />

Last F. M., Last W. M. and Halden N. M. (2010) Carbonate<br />

microbialites and hardgrounds from Manito Lake, an alkaline,<br />

hypersaline lake in the north Great Plains <strong>of</strong> Canada. Sed. Geol.<br />

225, 34–49.<br />

Menzel H. and Brückner A. (1930) Studien an kohlensauren<br />

Magnesiumsalzen. I. Basische Magnesiumcarbonate. Z. Elektrochem.<br />

36, 63–87.<br />

Morse J. W. and Casey W. H. (1988) Ostwald processes and<br />

mineral paragenesis in sediments. Am. J. Sci. 288, 537–560.<br />

Nashar B. (1965) Barringtonite – a new hydrous magnesium<br />

carbonate from Barrington Tops, New South Wales, Australia.<br />

Mineral. Mag. 34, 370–372.<br />

Ostwald W. (1897) Studien über die Bildung und Umwandlung<br />

fester Körper. Z. Phys. Chem. 22, 289–330.<br />

Power I. M., Wilson S. A., Thom J. M., Dipple G. M. and<br />

Southam G. (2007) Biologically induced mineralization <strong>of</strong><br />

dypingite by cyanobacteria from an alkaline wetland near Atlin,<br />

British Columbia, Canada. Geochem. Trans. 8, 13.<br />

Power I. M., Wilson S. A., Thom J. M., Dipple G. M., Gabites J.<br />

E. and Southam G. (2009) The hydromagnesite playas <strong>of</strong> Atlin,<br />

British Columbia, Canada: a biogeochemical model <strong>of</strong> CO2<br />

sequestration. Chem. Geol. 260, 286–300.<br />

<strong>Author's</strong> <strong>personal</strong> <strong>copy</strong><br />

Phase transitions in the system MgO–CO2–H2O 13<br />

Raade G. (1970) Dypingite, a new hydrous basic carbonate <strong>of</strong><br />

magnesium, from Norway. Am. Mineral. 55, 1457–1465.<br />

Russell M. J., Ingham J. K., Zedef V., Maktav D., Sumar F., Hall<br />

A. J. and Fallick A. E. (1999) Search for signs <strong>of</strong> ancient life on<br />

Mars: expectations from hydromagnesite microbialites, Salda<br />

Lake, Turkey. J. Geol. Soc. London 156, 869–888.<br />

Söhnel O. and Mullin J. W. (1978) A method for the determination<br />

<strong>of</strong> precipitation induction periods. J. Cryst. Growth 44, 377–<br />

382.<br />

Stamatakis M. G. (1995) Occurrence and genesis <strong>of</strong> huntite–<br />

hydromagnesite assemblages, Kozani, Greece: important new<br />

white fillers and extenders. Trans. Inst. Mining Metal 104,<br />

B179–185.<br />

Stephan G. W. and MacGillavry C. H. (1972) The crystal structure<br />

<strong>of</strong> nesquehonite, MgCO3 3H2O. Acta Crystallogr. B28, 1031–<br />

1033.<br />

Suzuki J. and Ito M. (1973) A new magnesium carbonate hydrate<br />

mineral, Mg5(CO3)4(OH)2 8H2O, from Yoshikawa, Aichi Prefecture,<br />

Japan. J. Jpn. Assoc. Mineral. Petrol. Econ. Geol. 68,<br />

353–361.<br />

Van Santen R. A. (1984) The Ostwald step rule. J. Phys. Chem. 88,<br />

5768–5769.<br />

Wells R. C. (1915) The solubility <strong>of</strong> magnesium carbonate in<br />

natural waters. Am. Chem. Soc. J. 37, 1704–1707.<br />

White W. B. (1971) Infrared characterisation <strong>of</strong> water and<br />

hydroxyl ion in the basic magnesium carbonate minerals. Am.<br />

Mineral. 56, 46–53.<br />

White W. B. (1974) The carbonate minerals. In Mineralogical<br />

Monograph 4. The Infrared Spectra <strong>of</strong> Minerals (ed V. C.<br />

Farmer). Mineralogical Society. pp. 227–282.<br />

Xiong Y. and Lord A. S. (2008) Experimental investigations <strong>of</strong> the<br />

reaction path in the MgO–CO2–H2O system in solutions with<br />

various ionic strengths, and their application to nuclear waste<br />

isolation. Appl. Geochem. 23, 1634–1639.<br />

Zeppenfeld K. (2006) Crystallization kinetics <strong>of</strong> strontianite from<br />

Sr(HCO3)2 solutions. Chem. Erde 66, 319–323.<br />

Zhang Z., Zheng Y., Ni Y., Liu Z., Chen J. and Liang X. (2006)<br />

Temperature and pH dependent morphology and FT-IR<br />

analysis <strong>of</strong> magnesium carbonate hydrates. J. Phys. Chem.<br />

110, 12969–12973.<br />

Associate editor: Alfonso Mucci

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

Saved successfully!

Ooh no, something went wrong!