Subscriber access provided by DALHOUSIE UNIV
Article
Controls of Microstructure and Chemical Reactivity on the Replacement of Limestone by Fluorite Studied Using Spatially Resolved Small Angle X-ray and Neutron Scattering Juliane Weber, Michael C. Cheshire, Victoria Distefano, Kenneth C Littrell, Jan Ilavsky, Markus Bleuel, Jessica Bozell-Messerschmidt, Anton V. Ievlev, Andrew G. Stack, and Lawrence M. Anovitz ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.9b00085 • Publication Date (Web): 29 Jul 2019 Downloaded from pubs.acs.org on August 5, 2019
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 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 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.
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 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Controls of Microstructure and Chemical Reactivity on the Replacement of Limestone by Fluorite Studied Using Spatially Resolved Small Angle X-ray and Neutron Scattering
Juliane Weber1*, Michael C. Cheshire1, Victoria H. Distefano1,2, Kenneth C. Littrell1, Jan Ilavsky3, Markus Bleuel4, Jessica K. Bozell-Messerschmidt5, Anton Ievlev6, Andrew G. Stack1, Lawrence A. Anovitz1 1Chemical
Science Division, Oak Ridge National Laboratory, Oak Ridge, USA. Current
Address: Lunar and Planetary Laboratory, University of Arizona, Tucson, USA. 2University
of Tennessee, Knoxville, USA.
3Argonne
National Laboratory, Chicago, USA.
4National
Institute for Standard and Technology (NIST), Gaithersburg, USA.
5University
6Center
of Nebraska, Omaha, USA.
for Nanophase Material Science, Oak Ridge National Laboratory, Oak Ridge,
USA.
This manuscript has been co-authored by UT-Battelle, LLC, under contract DE-AC05-00OR22725 with the US Department of Energy (DOE). The US government retains and the publisher, by accepting the article for publication, acknowledges that the US government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for US government purposes. DOE will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan). ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
*Corresponding
author:
[email protected],
[email protected] Abstract Fluid-mineral interactions can alter the pore structure and mineral composition of earth materials, sometimes leading to complete replacement of one mineral phase by another. A quantitative understanding of these processes is needed for the prediction of contaminant transport in nuclear waste management, oil and gas exploration, geothermal energy production and in many geological processes. Currently, a detailed understanding of how the original microstructure and chemical reactivity is affecting the replacement rate and porosity development is lacking, which would enable the prediction of contaminant transport.
Here, we present a systematic experimental study of limestone replacement in the model system calcite-fluorite varying both the texture and chemical reactivity of the parent rock. By combining X-ray ((U)SAXS) and neutron (ultra) small-angle scattering ((U)SANS) we quantified changes in the porosity as a function of depth within the sample and time. By
2
ACS Paragon Plus Environment
Page 2 of 101
Page 3 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
shielding the samples with a set of annular Cd-masks during neutron scattering, we obtained
spatially-resolved
porosity
information.
Microstructural
changes
were
investigated using scanning electron microscopy (SEM), and composition changes were assessed using chemical imaging by Time-of-Flight Secondary Ion Mass Spectrometry (ToF—SIMS) and SEM-Energy-Dispersive X-Ray Spectroscopy (SEM-EDX).
The replacement of limestone via fluorite takes place via advantageous pathways enhancing the available reactive surface area, e.g. fractures, accessible, interconnected porosity, non-connected porosity grain boundaries and twin boundaries. Our results presented here emphasize the importance of a detailed structural and textural assessment of the starting material both in experimental studies and in modelling studies of natural processes to make accurate predictions about reaction rates.
Keywords: Small angle scattering, Small angle neutron scattering, Small angle X-ray scattering porosity analyses, dissolution-reprecipitation, replacement, multi-scale analyses
3
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
1
Page 4 of 101
Introduction
Fluid flow in sedimentary rocks controls the migration and retention of water, gas, hydrocarbons and contaminants, as well as cement precipitation, alteration and replacement of the host mineralogy, and many other diagenetic processes. These properties, in turn, are controlled by the microstructure and evolution of pore space in geologic formations, in which the size, distribution and connectivity of the pores dictate fluid transport and surface wetting. Furthermore, accessible porosity, which includes grain boundaries, defines the reactive surface area of the mineral-water interface and, therefore, controls the rates of replacement reaction between solid mineral phases, which are among the most complex mineral reactions in nature.1–3
When fluids interact with rocks, the mineral phases of which the rock consists can be replaced by another mineral phase. These replacement reactions can proceed via: (1) solid-state
diffusion4,5
and
(2)
dissolution-reprecipitation1,6.
In
both
exchange
mechanisms, the parent mineral may be replaced in-situ by the product, often retaining its original morphology, forming pseudomorphs. Although, dissolution-reprecipitation, 4
ACS Paragon Plus Environment
Page 5 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
transport and reprecipitation at a different site are also common. In most cases, diffusion is too slow at ambient conditions to be a relevant exchange mechanism5, although it is a
well-known phenomenon at higher temperatures.7–12 Recent studies indicate that
dissolution-reprecipitation may be the governing mechanism for such replacement reactions at lower temperatures.1,6 These reactions proceed sequentially via growth of a
boundary fluid layer at the fluid-mineral interface. By dissolution of the surrounding phase, the fluid boundary layer can become supersaturated with respect to a new phase, i.e., the product phase, whose growth or precipitation can then occur. In a multi-grain rock, however, this process is further limited by transport to the reaction site. Reacted zones may, therefore, either form an outer “skin” on the rock or be more diffuse depending on the significance and rate of grain boundary transport.13–17 In recent years, it has been
recognized that such replacement reactions are ubiquitous in geologic systems and also impact a number of energy-related industrial activities, e.g. mining and nuclear waste storage.3,18,19
5
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
One of the most important characteristics of replacement reactions is the change in the porosity associated with them. In early studies, it was assumed that the change during these replacement reactions is defined only by the difference in molar volume between the primary and the secondary phases. Since then, however, several studies have identified that changes in porosity are due to a combination of molar volume and solubility differences.20–24 In order to both predict how geological systems evolve and
examine the reactive history of a rock, therefore, the physical and chemical “fingerprints” of these replacement reactions must be explored using more realistic systems. This begins by investigating replacement reactions in rock samples, expanding on previously published results of laboratory experiments using single crystals.6,20,25–29 Few previous
studies on the effect of the microstructure on replacement reactions have been published,30–32 which have indicated the importance of the initial microstructure and its effect on the replacement reaction. Because of the existence of grain boundaries and other pre-existing pore structures, porosity changes in real rocks span a wide range of length scales from nanometer to macroscopic. This makes making complete porosity 6
ACS Paragon Plus Environment
Page 6 of 101
Page 7 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
characterization challenging, and the overall nature of the transformation more complex.33
In theory, two factors control porosity generation: (1) the difference in the molar volumes of the parent and the product phases, and (2) their relative solubilities at the mineral-water interface, which in turn may be affected by changes in fluid properties under spatial confinement.2,20,34,35 Volume-reducing reactions, with high molar volume differences between the parent (higher molar volume) and the product (lower molar volume) will lead to material loss, and subsequent increased porosity [in percent] as:
Δ𝑉 = 100 ×
(
𝑛𝑝𝑉𝑚,𝑝 ― 𝑛𝑑𝑉𝑚,𝑑 𝑛𝑑𝑉𝑚,𝑑
)
(1)
Where np and nd are the number of moles precipitated and parent dissolved and Vm,p and Vm,d are the molar volumes of the precipitating and dissolving phases, respectively. As the solubility of parent and product phase control the number of moles precipitated and dissolved, this equation also accounts for the solubility differences between parent and product phase.20 Volume-increasing replacement reactions,
7
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
however, lead to pore sealing, fractures and/or pressure-induced crystallization.6,36–38
More accurate estimations of the expected volume changes must also account for the solubility ratio of the two minerals, as the fluid volumes available for exchange are usually limited.20 However, it is not always clearly observed that volume-decreasing reactions
lead to porosity increase and vice-versa, as previous studies have shown porosity generation during replacement processes that involve both molar volume decreases as well as increases.36–39 As noted above, the latter may reflect the formation of multiscale
fracture networks generated by the pressure of crystallization.
A number of studies have considered porosity changes during dissolution reprecipitation reactions,1,24,36–38,40 and furthermore, several studies have investigated
the effect of pre-existing microstructure on replacement reactions.31,32,41,42 Naturallyreacted samples obtained during field studies on the other hand, contain significant microstructure43–46 but are often too complex to evaluate completely as the conditions
under which they formed are, at best, uncertain. We have, therefore, performed controlled 8
ACS Paragon Plus Environment
Page 8 of 101
Page 9 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
laboratory replacement experiments on samples from three rock formations, two limestones and a dolostone, to study the relations between microstructure and replacement extent and speed.
In this study, we varied the microstructure of the starting material to investigate the effect of that microstructure on the replacement reaction. To do so, we replaced the calcite and/or dolomite in the starting material with fluorite by reaction with a saturated ammonium fluoride solution as:
CaCO3 + 2NH4F CaF2 + 2NH3 + H2CO3
(2)
The expected volume change based on the molar volume of parent and product phase for this reaction, (Δ𝑉) is -33.51 vol. %. Thus, significant porosity should be formed. A previous study by Pedrosa et al.47 conducted replacement experiments of calcite by fluorite and included the product and parent phase solubilities in this equation, leading to an expected porosity change of (Δ𝑉) = 30 vol %.
The equation for the reaction of dolomite with saturated ammonium fluoride solution is:
9
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
CaMg(CO3)2 + 4NH4F CaF2 + MgF2 + 4NH3 + 2H2CO3
Page 10 of 101
(3)
The change in molar volume for this reaction (Δ𝑉 = -32.43 %) is very similar to that of calcite-fluorite replacement (Reaction 2).
Reaction (2) was previously utilized by Pedrosa et al.47 to study the replacement
of calcite single crystals by fluorite. In these experiments, replacement of 1 mm cubes was observed after several hours at temperatures between 60 and 140 °C.47 Despite the
large change in molar volume involved, within the accuracy and resolution of SEM image analyses, the original shapes and volumes of the calcite crystals were preserved. The product fluorite was needle-shaped and oriented perpendicular to the reaction interface, which created a high porosity within the replaced sample. A distinct densification of the older (outer) fluorite was observed, along with the more fractured younger (inner) replaced region. The origin of this difference is uncertain. In a second study, Carrera marble was replaced by fluorite experimentally demonstrating a log-linear (power law)
10
ACS Paragon Plus Environment
Page 11 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
correlation between surface area, porosity and amount of replaced calcite with reaction time.48
Because previous studies indicated complex changes in the porosity during the replacement reactions, we have used spatially resolved multi-scale characterization methods for our porosity analyses. To span a wide range of pore sizes, we combined small angle scattering with ultra-small angle scattering using both neutron and X-rays, yielding spatially-resolved information on pore geometries at scales from 100 nm to 30 µm. By varying the initial porosity and chemical reactivity of the pristine limestone, we aimed to determine how the microstructure of the starting material influences replacement rates, extent and porosity changes during reaction. The relative effects of chemical reactivity and initial microstructure, however, probably depend on the magnitude of each. A reduced chemical reactivity in highly permeable rocks will likely lead to a diffuse reaction zone, while the opposite is likely to form a well-differentiated rim.
11
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
2
Materials and Methods
2.1.
Characterization of Starting Materials
To test how initial microstructure affects replacement, we varied the microstructure of the limestones in our experiments by selecting two limestone of different porosity (high porosity Texas Cream “TC” limestone and low porosity Carthage Marble “CM” limestone), and a low porosity dolostone (Wisconsin Dolomite, “WD”). The porosity of the WD dolostone is comparable to that of the CM limestone but its reactivity is expected to be lower as solubility of dolomite (logKsp = -17.0949) is significantly lower than that of calcite
(logKsp = -8.4849). Furthermore, growth experiments using single crystal calcite have
demonstrated the inhibiting effect of Mg2+ in solution on calcite growth50,51. Replacement
experiments with NH4F were conducted using three different rock types: Carthage Marble (“CM”), Texas Cream Limestone (“TC”) and Wisconsin Dolomite (“WD”). These rocks were chosen for their relative pure composition and availability. Powder X-ray diffraction patterns were modeled via Riedveld refinement to obtain mineralogical composition of the starting material. Carthage marble is 100 wt% calcite with trace quantities of quartz. 12
ACS Paragon Plus Environment
Page 12 of 101
Page 13 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Texas Cream limestone is 99 wt% calcite, 0.5 wt% sylvite and 0.5 wt% quartz. Wisconsin dolomite is 80 wt% dolomite, 18 wt% quartz, 2 wt% orthoclase and contains trace amounts of kaolinite. Porosity of the original material varies, with TC being the most porous material with approximately 25% porosity, WD with 3% and CM with 1.5% based fluid saturation porosity determination (see Supplementary Information, Table S1). SEMBSE images of the starting materials are given in Figure 1.
13
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
14
ACS Paragon Plus Environment
Page 14 of 101
Page 15 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Figure 1
Scanning Electron Microscopy Backscattered Electron (SEM-BSE) contrast
images of the starting materials. (a) High porosity TC limestone, (b) low porosity CM limestone and (c) low porosity WD dolostone. The low porosity WD dolostone contains dolomite (dol), orthoclase (or) and quartz (qtz) as identified by XRD and SEM-EDS mappings (see Supporting Information, Figure S1 for mappings.)
The average porosity of our starting material was analyzed by fluid saturation porosity determination and (U)SANS. Both porosity determinations are in good agreement for CM and WD limestone, with 1.5% (1.8% by (U)SANS) for CM and 3% (2.2% by (U)SANS) for WD. For the high porosity TC limestone, the discrepancy between the two methods is larger with 25% (8% by (U)SANS). Most likely, this is caused by a significantly higher amount of macropores for the TC in comparison to the other two rocks. This is suggested by the Pore Size Distribution (PSD) analyses of the (U)SANS data (see Supplementary Information for PSD graph, Figure S2). SEM characterization of TC also showed a high percentage of macropores, often in the complex shapes of calciferous fossils (Figure 9 a). As (U)SANS characterization only probes pores between 100 nm and 30 µm, (U)SANS systematically underestimates the total porosity of the TC material.
15
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
2.2.
Replacement Experiments
Samples of all three rock types were cut into cylinders 1.5875 cm long by 1.5875 cm diameter and reacted in steel pressure vessels with Teflon liners and a Teflon seal.52 Teflon liner dimensions are 1.905 cm inner diameter and 2.54 cm height with a
fluid volume of 12 ml. Experiments were fluid saturated and sealed to avoid water loss during the reaction. Cores were weighed before and after the replacement experiments. After the rock was placed in the vessel, it was packed with extra NH4F salt to ensure that the fluid remained saturated during the replacement reaction, and saturated NH4F solution was then added. Due to the additional salt in the reaction vessel, the solid/liquid ratio was slightly different in each experiment. For all conducted experiments, the solid/liquid ratio was between 316 and 525 g/L. Vessels were then sealed, preheated in boiling water to minimize the time needed to heat the vessel to the reaction temperature, and then placed in an aluminum block in a convection oven at 140 °C. An overview of the experiments conducted is given in Table 1.
16
ACS Paragon Plus Environment
Page 16 of 101
Page 17 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
After removal from the furnace, the vessels were quenched in running cold water, opened, and the rock core was left in DI water overnight and subsequently dried. For further analyses, polished thin sections 150 µm thick were prepared by Spectrum Petrographics of each sample mounted on quartz glass slides (cf. Anovitz et al., 200953).
Table 1 Overview of experiments conducted in this study.
Experiment name
Rock type
TC_Reference
TC
CM_Reference
CM
WD_Reference
WD
TCF2
TC
CMF2
CM
WDF2
WD
TCF4
TC
CMF4
CM
WDF4
WD
TCF8
TC
CMF8
CM
WDF8
WD
TCF22
TC
CMF22
CM
WDF22
WD
TCF45
TC
CMF45
CM
WDF45
WD 17
ACS Paragon Plus Environment
Duration [days] -
2
4
8
22
45
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
2.3.
Optical Microscopy Analyses
Thin-sections were imaged using an Olympus BX60 microscope outfitted with a Nikon D700 camera in reflected light to obtain initial examinations of the alteration in each experiment and identify suitable regions for SEM characterization.
2.4.
Scanning Electron Microscopy (SEM) Characterization
Low-magnification (~300x), two-dimensional (2D) SEM/BSE images of unreacted and reacted cores were obtained from the polished thin sections using a Hitachi S3400 environmental scanning electron microscope operated at 50 Pa vacuum pressure (to eliminate charging of insulating samples) using a 15 kV electron beam. The 300x fold magnification provided a pixel edge length of 661 nm for 640 x480 pixel image, yielding an image size of 423.0 x 317.3 µm. In addition, samples were also imaged using a Hitachi S4800 SEM instrument operated at 15 kV electron beam in high-vacuum mode for higherresolution imaging using different magnifications. For imaging on this instrument, the thin
18
ACS Paragon Plus Environment
Page 18 of 101
Page 19 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
sections were coated with a ~30 nm of carbon using a Crossington 208 carbon coater (TedPella Inc., USA). Reaction rim thicknesses were measured using ImageJ.
2.5.
X-ray Scattering Analyses
Thin sections were analyzed via ultra-small (USAXS) and small angle X-ray scattering (SAXS) for pore distributions and wide-angle X-Ray scattering (WAXS) (to reveal Bragg peaks of calcite/dolomite and fluorite phases) at the Advanced Photon Source at Argonne National Laboratory on beamline 9ID with 21 keV X-rays.54 This
instrument is a Bonse-Hart USAXS instrument, equipped with SAXS and WAXS pinhole cameras. Data were corrected for empty beam scattering and reduced using instrumentprovided data correction routines.55 Data were desmeared. The design of the instrument
guarantees that the Q ranges of different segments overlap. Collected data were put on an absolute intensity scale to enable quantitative porosity characterization. The beam size was 0.5 mm for (U)SAXS measurements. For (U)SAXS measurements, a line scan of 13 – 14 measurement points with 1 mm between them was scanned along the diameter
19
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
of the samples core. For SAXS and WAXS, beam size was 0.2 x 0.5 mm, with data collection time of about 4 minutes per data set.
2.6.
(Ultra) Small Angle Neutron Characterization
The analytical approach for the (U)SANS characterization is described in detail in our previous papers33,53,56–60 and is therefore only summarized here briefly. Previous
studies demonstrated that pore structures of geological materials in the size range from ~200 nm to ~20 µm can be characterized using small- and ultra-small angle neutron scattering ((U)SANS).33,53,59,61,62 USANS measurements were conducted at the BT5
USANS instrument at the NIST Center for Neutron Research (NCNR)63 and the BL-1A
TOF-USANS instrument at the Oak Ridge National Laboratory (ORNL) Spallation Neutron Source (SNS) in Oak Ridge, USA.64 The wavelength of neutrons was 3.6 Å and the width of the Darwin plateau was 5.3 arcsec. For SANS measurements, there were three instrument configurations (sample-to-detector distance of 1 and 7 m with λ = 4.75 Å, respectively, and 19m with λ = 12 Å). The resultant scattering vector ranged from 6 x 10-5 to 3 x 10-3 Å-1, which corresponds to pore diameters of 100 nm and 5 µm. 20
ACS Paragon Plus Environment
Page 20 of 101
Page 21 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
For the BT5 USANS instrument at NCNR, the wavelength of the neutrons was 2.38 Å and the width of the Darwin plateau was 5 x 10-5 Å-1. The resultant scattering vector ranged from 3 x 10-5 to 2 x 10-3 Å-1, which corresponds to pore diameters of 300 nm and 30 µm. In both cases USANS data were corrected for empty-beam scattering, background counts, sample transmission and scattering volume and reduced to an absolute scale (differential cross-section) to the intensity of the direct beam using IGOR Pro.65
SANS measurements were performed on the NGB 30 m instrument66 at the
National Institute for Standard and Technology (NIST) Center for Neutron Research (NCNR) and the GP-SANS67 at the High Flux Isotope Reactor (HFIR) at ORNL, TN. On
the NCNR instrument, four sample-to-detector distances (1, 4 and 13 m, and 13 m with lenses) were used with λ = 6 Å and 8.4 Å at 13 m. This ensured an extension of the Q range to lower values and a better overlap of the USANS data at low q. The same sampleto-detector distances and wavelengths were used on the HFIR instrument. Data reduction
21
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
and desmearing were performed using data reduction software provided by NIST/NCNR65 and SNS/HFIR.67
The physical density and chemical composition of a material determines its scattering length density. In turn, the scattering intensity from an interface is a function of the square of the difference in the scattering length densities of the two phases (an empty pore is considered a “phase” in this cases) across that interface and the number of such interfaces of a given size. For a rock, the average scattering length density therefore depends on the volume percentage of each mineral it contains. X-ray and neutron scattering densities were determined based on the XRD analyses and are given in Table 2. For the determination of sample porosity from scattering curves, a two-phase system was assumed with the majority of scattering occurring at the mineral/pore interface.9,62
22
ACS Paragon Plus Environment
Page 22 of 101
Page 23 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Table 2 Scattering length densities used in this study.
Starting
Wisconsin
Carthage Marble
Texas Cream (TC) Fluorite
Material
Dolomite (WD)
100% calcite (CaCO3)
99% calcite
(CaF2)
80% dolomite
(CaCO3),
(Ca,Mg(CO3)2),
0.5% sylvite (KCl)
18% quartz
0.5 wt%
(SiO2),
(SiO2)
quartz
2% orthoclase (KAlSiO4)
X-ray
23.70
22.97
22.93
scattering length density [1010cm-2]
23
ACS Paragon Plus Environment
26.26
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Neutron
5.16
4.72
Page 24 of 101
4.71
4.72
scattering length density [1010cm-2]
A special set of cadmium masks were fabricated for these experiments to analyze radial changes in porosity. These are shown in Figure 2, and described in Table 3. Except for the innermost, mask 0, these consisted of rings cut in 0.5 mm thick Cd sheet, with the inner part supported by thin Cd bars. Cd of this thickness is sufficient to block neutrons of the energies used here, and the masks were attached directly to the front of the sample, so the only part of the sample from which scattering could occur was that exposed by the slits in each mask. The innermost edge of each ring was the same as the outermost edge of the next smallest, so all the available area was covered, and the widths of each ring was selected to be reasonably simple to machine while yielding nearly equal areas for 24
ACS Paragon Plus Environment
Page 25 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
each. Absolute intensities are corrected for actual sample areas, so the slight differences between the areas of the masks are immaterial. Each mask also had a series of scribed marks aligned to its center. During the analyses the largest mask was used first, and this was aligned to the outer edge of the sample. A fine pen was then used to mark the positions of the scribed lines onto the glass slide, so that subsequent, smaller, masks could be aligned to the same center on the sample. As analytical time was limited, the full mask suite was only measured for some of the experiments. Only mask 0 (core) and 4 (rim) were analyzed for the shorter runs.
Figure 2
Radial Cd masks designed for this experiment.
25
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Table 3 Cd mask geometries.
Mask # 0 1 2 3 4
Outer Diameter [cm] 0.71438 1.03187 1.27 1.42875 1.58750
2.7.
Inner Diameter [cm]
Slit width [cm]
Area [cm2]
0.71438 1.03187 1.27 1.42875
0.71438 0.15875 0.11906 0.07938 0.07938
0.15773 0.17145 0.16942 0.13259 0.14808
Blocked Area [cm2] 0 0.15773 0.32918 0.49860 0.63119
ToF-SIMS Characterization
Time-of-Flight Secondary Ion Mass Spectrometry (ToF-SIMS) measurements were carried out using a TOF.SIMS.5.NSC instrument (ION.TOF GmbH, Germany). A bismuth (Bi3+) liquid metal ion gun with an energy of 30 keV, current of 0.5 nA and spot size ~120 nm was used as a primary source for chemical imaging. Scanning was performed over areas of 250 x 250 µm with a resolution of 512 x 512 pixels yielding a pixel size of 488 nm. Analysis of the secondary ions was performed using time-of-flight mass analyzer run in both positive and negative ion modes with a mass resolution Dm/m = 100 — 200. Charge compensation was carried out using a low energy electron flood gun. 26
ACS Paragon Plus Environment
Page 26 of 101
Page 27 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
2.8.
Small Angle Scattering Data Analyses
Both X-ray and neutron scattering data were analyzed using the IRENA and NIKA macro packages for IgorPro.55 Pore size distributions were calculated in Irena using the
total non-negative least square (TNNLS) method, which implements the work of Merritt and Zhang.55,68–70 In this method, a model pore size distribution of spheroid particles
(aspect ratio 1) are fitted to the scattering data. The minimum diameter was set to 100 nm and the maximum diameter to about 30 µm. Uncertainties calculated in Irena were obtained by running multiple fits to the data, and varying the data by adding Gaussian noise.55 A detailed description of how to apply scattering to the characterization of
porosity and pore structures can be found in Anovitz and Cole.33
3
Results
3.1.
Effect of the Microstructure on the Replacement of Limestone by Fluorite
Partial to full replacement of the limestone was observed (see Supporting Information, Figures S3, S4 and S5), depending on the starting microstructure of the pristine rock and reaction time. In general, the rock cores maintained their original shape 27
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
during the replacement within the error of analyses (several µm in SEM characterization). On some of the limestone cores, however, it was observed that the outermost layer became very brittle and was subject to disintegration when the core was dried, indicating significant microstructural changes due to replacement.
Changes in weight before and after replacement were used to get an initial, rough estimate of the extent of replacement (see Supporting Information, Table S2, Figure S6). This would, of course, be affected by the sample loss mentioned above. The high porosity TC limestone, after an initial decrease in weight, showed a rather constant decrease in weight until the end of the experiment. In comparison, the low porosity CM limestone showed a rather steep decrease in weight followed by an increase (see Supporting Information for a more detailed analyses). Due to sample embrittlement and subsequent loss, there is a significant error associated with these weight determinations.
As noted above, (U)SANS analyses of the cross-sections of the rock cores were performed using annular Cd-masks to shield different radii of the samples, which allowed us to spatially resolve and quantify the porosity generation as a function of depth within 28
ACS Paragon Plus Environment
Page 28 of 101
Page 29 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
each core. These measurement techniques can be used to determine the average porosity within the material volume exposed to the neutron beam by the Cd-shielding. (U)SANS data (Figure 3 a) indicate increased porosity in the replaced rim after two days of reaction with NH4F for both the low and high porosity limestones. The resulting rim porosities for the two limestones are similar: 14.7 % for the low porosity CM limestone and 14.4 % for the high porosity TC limestone, as compared to 1.8 and 8 % porosities for the unreacted materials, respectively.
In both cases, however, the total (U)SANS porosity (14.7 % for the low porosity CM limestone and 14.4 % for the high porosity TC limestone) is less than would be expected from total fluorite replacement of a solid calcite crystal (30% replacement expected when considering both molar volume and solubilities47). Possible reasons for this could be a preexisting porosity, incomplete replacement reactions or the generation of macropores that are larger than what (U)SANS measurements can detect. Furthermore, the TC limestones contained 0.5 wt% quartz and 0.5 wt% sylvite, which is a larger amount of impurities than average for this type of sample (99.97 wt% CaCO3 with 29
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
0.02 wt% Mg and 0.01 wt% Sr for the islandic spar and 99.7 wt% CaCO3 with 0.03 wt% Mg for Carrera marble) contained in calcite samples used in previous experiments by Pedrosa et al.47,48. The porosity as a function of distance from the center of the core in the two-day experiments also increases approximately linearly with depth for both CM and TC (Figure 3). However, the relative porosity changes of 711% for the low porosity CM limestone is more than an order of magnitude higher than for the high porosity TC limestone (73%). This indicates that the porosity of the product phases, fluorite, is similar in the two reactions and therefore independent from the porosity of the parent phase.
A distinctly different porosity structure appeared after 45 days of reaction depending on the microstructure of the starting limestone (Figure 3 b). (U)SANS characterization of the low porosity CM limestone indicates an altered zone with up to 20% porosity, surrounded by an outermost rim of only ~13 % porosity. The inner core is still unreacted and shows a low porosity of 1-2 %, which is consistent with our measurement of the original CM limestone (dotted line in Figure 3 b). In the case of the 30
ACS Paragon Plus Environment
Page 30 of 101
Page 31 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
high porosity TC limestone, however, the measured porosity at all depths within the core is higher than that of the parent phase indicating porosity generation and reaction throughout the core.
Figure 3
(U)SANS characterization after 2 and 45 days. (a) (U)SANS analyses of average
porosity in distance from the center of the core for low porosity CM and high porosity TC limestone reacted for 2 days. Porosity increases at the rim to 14 % for both limestones. (b) (U)SANS analyses of the porosity in limestones reacted for 45 days showing a reduction of porosity in the outer rim of both high and low porosity limestone.
The decrease in porosity near the rim observed in the (U)SANS characterization of the CM limestone sample after 45 days indicates that the replacement reaction is
31
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
accompanied by a secondary textural re-equilibration that is the final step of the replacement reaction.29 That is, after the initial replacement, a secondary, textural reequilibration occurs in which porosity is decreased by recrystallization (Figure 4).
Figure 4
SEM characterization of CM limestone replaced by fluorite after (a) 8, (b) 22 and
(c) 45 days. Near the outer rim of the samples, a region of reduced porosity is visible (marked by 1). At the rim between limestone (Calc) and fluorite (Fl), acicular fluorite growth is visible (2).
32
ACS Paragon Plus Environment
Page 32 of 101
Page 33 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Additional (U)SANS analyses (Figure 5 a) of samples of the low porosity CM limestone reacted for 2, 8, 22 and 45 days support our hypothesis that the replacement reaction is initially accompanied by an overall increase in porosity relative to the unreacted material before a decrease of porosity in the outermost rim can be observed. These secondary replacement layers can be clearly seen in the near-rim areas of Figure 4. A similar secondary replacement layer can be observed, but was not commented on, in the work of Pedrosa et al., 2017.48 Porosity at the center of the rock core remains constant
indicating that the replacement does not proceed throughout the complete rock core. Similarly, we also observed a decrease in porosity of the outermost layer for the higher porosity TC limestone, with reaction time again indicating a secondary recrystallization similar to that observed for the low porosity CM limestone (Figure 5 b). In contrast to the CM limestone, however, the TC limestone was replaced completely before the secondary replacement began. Therefore, textural re-equilibration seems to be independent of the extent of replacement. An interpretation consistent with our (U)SANS results would be, assuming a constant supply of reactant, that the rate of replacement with depth is 33
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
primarily a function of the porosity/permeability, and reactive surface area of the limestone.
Figure 5
Porosity distribution determined by (U)SANS for (a) low porosity CM limestone and
(b) high porosity TC limestone reacted for 2, 8, 22 and 45 days.
(U)SANS analyses were augmented with (U)SAXS analysis, which provides porosity analyses in a similar pore size range but with higher spatial resolution as the beam footprint on the sample, and therefore the sample volume generating the signal, is smaller. In addition, simultaneous wide-angle X-ray scattering (WAXS) can provide phase identification for each point analyzed. Here, we conducted line-scans from one edge of
34
ACS Paragon Plus Environment
Page 34 of 101
Page 35 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
the core to the other consisting of 13 - 14 individual measurement points with (U)SAXS and WAXS spectra collected at each point. WAXS data were used to identify the sampled mineral phase (see Supporting Information, Figures S7 - S13 and Tables S3 - S11). Samples from day 4, 8 and 22 were characterized using (U)SAXS and WAXS.
35
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Figure 6
: Porosity changes in CM limestone with increasing reaction time (a) 4 days reaction time,
(b) 8 days reaction time and (c) 22 days reaction time. Diagrams show porosity [%], Mean and Median Diameter [µm] for each (U)SAXS measurement point collected during linescan traversing the middle of the core indicated in (a), (b) and (c). Color code for measurement points indicates phases identified by WAXS pattern (for individual WAXS pattern see Supporting Information, Figures S7 - S9 and Tables S3 - S5), with orange indicating calcite, red indicating fluorite and pale yellow indicating a calcite/fluorite mix. Note the recrystallized area near the rim after 22 days.
For the high porosity TC limestone (Figure 7), (U)SAXS analyses show a uniform porosity. After four days of reaction time, the high-porosity TC limestone was completely
36
ACS Paragon Plus Environment
Page 36 of 101
Page 37 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
replaced by fluorite as indicated by the WAXS diffraction pattern (see Supplementary Information). Once the limestone is replaced by fluorite, the (U)SAXS data also show no changes in the microporosity (pore sizes of 200 nm – 20 µm), which remains constant at around 10 % with a uniform distribution throughout the cross-section of the core based on (U)SAXS analyses (Figure 7) – only about 2 % higher than the porosity of the starting material. The mean and median pore sizes are approximately constant throughout the core as well (for individual values see Supporting Information, Tables S6 - 8, Figure S14), although there is some suggestion that it decreases near the rim. (U)SANS analyses, however, indicated a higher porosity in the rim of ~20 % averaging over a larger area, which can possibly be explained by the fact that the (U)SANS averages over a larger area than the highly localized (U)SAXS analyses.
37
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Figure 7
Porosity changes in TC limestone with increasing reaction time (a) 4 days reaction
time, (b) 8 days reaction time and (c) 22 days reaction time. Diagrams show porosity [%] on the left y-axis and on the right y-axis mean and median diameter [µm] for each (U)SAXS measurement point collected during linescan traversing the middle of the core. The orange color code indicates that the measurement points were identified in WAXS pattern as fluorite (see Supporting Information, Figures S12 and S13, Tables S6 - S8 for individual patterns). Again, note the recrystallization of the outer edge.
The pore size distributions (PSD) of the low-porosity CM limestone generated from the (U)SAXS cross sections (Figure 8) further strengthen our hypothesis that, for a low 38
ACS Paragon Plus Environment
Page 38 of 101
Page 39 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
porosity starting material, replacement results in generation of new nano-scale porosity. The PSD of the pristine low-porosity CM limestone is significantly narrower than that of the newly precipitated fluorite (Figure 8). For material reacted for 4, 8 and 22 days, we can clearly see a shift towards higher total porosity, a wider pore-size distribution, and an increase in nanoporosity in comparison to the unreacted core. The PSD of the newly precipitated fluorite also indicates an increase in nanopores consistent with the observed decrease in the mean/median pore diameter (see Supporting Information, Figures S14 and S15). For the high porosity TC limestone, the initial PSD is broader and more macropores are visible. As the high porosity TC limestone was replaced faster, it is, however, possible that the nanoporosity coalesced and formed new macropores. PSD graphs for high-porosity TC limestone are given in the Supporting Information (Figure S16).
39
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Figure 8
: Pore Size Distribution analyzed by (U)SAXS for low porosity CM limestone replaced with
fluorite for 4 days (a), 8 days (b) and 22 days (c). Sample regions probed by (U)SAXS, which are fluorite as indicated by WAXS, are depicted in red, calcite is green and a mixture of the two phases is blue. Sample locations are the same as in Figure 7.
40
ACS Paragon Plus Environment
Page 40 of 101
Page 41 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
To chemically identify the replacement phase(s), we coupled the small angle scattering data with scanning electron microscopy (SEM) with energy-dispersive X-ray spectroscopy (EDX). For both high- and low-porosity limestone (TC and CM), a clear reaction front between a reacted rim and an unreacted pristine rock in the middle of the core was observed in samples reacted for two days (Figure 9). For the low-porosity limestone (CM), the reacted rim was visible for all reacted samples in good agreement with (U)SAXS measurements and, as noted above, a secondary recrystallized rim was visible (Figure 4). The overall reaction front observed via SEM in the low-porosity CM limestone looks similar after 8, 22 and 45 days. Higher resolution SEM images (Figure 9) show that the reaction at the interface proceeds faster along cracks and grain boundaries, which thereby provide a fast pathway for fluids into the low porosity rock. Our data demonstrates, therefore, that, at least for the low-porosity/permeability limestone replacement by fluorite proceeds in a manner similar to that previously reported for dolomitization,71,72 in which grain boundaries act as reactive pathways.
41
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
When comparing BSE-contrast images to EDX-mapping for low porosity CM limestone reacted for 45 days (see Figure 10), we can clearly see a higher concentration of silicon at the grain boundary region. The origin of this silicon is unknown, but it probably indicates that smaller silicon-containing phases, e.g. micas, clays or quartz grains, are being dissolved during the experiment, releasing silicon, which then moves along the grain boundaries and reprecipitates, possibly during quenching. Its presence is, however, further proof of the importance of grain-boundary transport in the replacement process.
The newly formed fluorite crystals in the low porosity CM limestone exhibit acicular growth, but the individual crystals appear to be randomly oriented (Figure 4). In the prior study of Pedrosa et al.,47 in the case of pure calcite replacement, formation of oriented,
sheet-acicular crystals was reported, similar to other reports of the replacement of calcite by strontianite.41 In our experiments, the fluorite needles are randomly oriented, which
could be caused by the space that is available for precipitation within the limestone. Possibly, this acicular orientation is caused by the replacement, which proceeds faster
42
ACS Paragon Plus Environment
Page 42 of 101
Page 43 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
along the grain boundaries than within the grains (Figure 11 b and Figure S17 of Supporting Information).
43
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
44
ACS Paragon Plus Environment
Page 44 of 101
Page 45 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Figure 9
: BSE-SEM images of high porosity TC and low porosity CM limestone at day 0, 2 and 45
of replacement reaction. The high porosity TC limestone (a) shows a reaction rim after 2 days of replacement (c) and is completely replaced after 45 days (e). Higher porosity in the replaced region is visible. The low porosity CM limestone (b) shows a clear reaction front indicating enhanced replacement along the grain boundaries after 2 days (d) and after 45 days (f).
Figure 10
SEM and SEM-EDX characterization of the replacement for the low porosity CM limestone.
(a) BSE image of low porosity CM limestone replaced for 45 days (CMF45) (b) Si distribution (c) Ca distribution and (d) F distribution. Enhanced F incorporation along the grain boundaries is visible.
For the high-porosity TC limestone, a reaction rim is visible after 2 days of reaction, which separates the unreacted limestone parent phase from the product phase (Figure 12). In comparison with CM, the higher porosity and permeability of the TC limestone increases 45
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
the surface area available for replacement. Therefore, the transport through the open matrix pore structure is more important than transport along grain boundaries. Thus, the sample becomes rapidly replaced in a more wholesale manner, leaving only larger grain cores to be replaced with time. Interestingly, we also do not observe acicular growth of the fluorite crystals in the product phase of the high porosity limestone experiments. This suggests that due to the higher porosity in the TC limestone, the new fluorite phase had more space during crystallization allowing an idiomorphic crystal growth (Figure 11).
46
ACS Paragon Plus Environment
Page 46 of 101
Page 47 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Figure 11
Two types of interface between original phase and product phase imaged SEM-BSE
contrast images of low porosity limestone reacted for 2 days (CMF2). (a) Interface with a gap between parent and product phase (b) Acicular fluorite crystal growth along preferential pathways.
47
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Figure 12
Chemical mapping of the interface between limestone and fluorite in TCF2 sample showing
uniform incorporation of fluorine into limestone. The higher porosity of fluorite-replaced regions is visible in (a) with rectangular mapping area (b) marked. Mappings for Si (c), F (d) and Ca (e) show incorporation of fluorine into the limestone.
Our experimental results clearly indicate a faster replacement reaction for the limestone with a higher starting porosity. This strengthens the hypothesis that porosity acts as a fast pathway for fluorite solution to reach the inside of the rock core. Furthermore, the higher porosity provides more surface area which accelerates the replacement reaction.
3.2.
Effect of the Reactivity on Replacement by Fluorite
To identify the effect of chemical reactivity of the replacement reaction, we also reacted Wisconsin dolostone (WD) with the saturated NH4F solution. The starting porosity (3%
48
ACS Paragon Plus Environment
Page 48 of 101
Page 49 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
based on fluid saturation analysis) of the low porosity WD dolostone is similar to the low porosity limestone CM (1.5% porosity). Like the limestone reaction, there was a tendency for reacted material to spall off the exterior of the cores. Gravimetric characterization after reaction with NH4F indicates relative weight decrease of up to 10%, similar to that observed for the low porosity CM limestone (Supporting Information, Table S2 and Figure S6).
(U)SANS observations (Figure 13 a) again show an increase in porosity in the outermost layer of the dolostone after replacement similar as observed above for limestone replacement. After two days of reaction with NH4F, an average porosity of 18.5 % is observed in the outermost rim, which is higher than the observed increase for both TC and CM limestone. After 45 days, porosity determined by (U)SANS is increased for all measurements in comparison to the pristine material with 3.9, 2.9, 7.8 and 4.2 % porosity for 0.4 cm, 0.5 cm, 0.6 cm and 0.8 cm distance from the core, respectively. Similar to the pure limestones, the dolostone also exhibits a secondary, textural reequilibration step in which porosity is reduced in the outermost rim (Figure 13 b) 49
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Figure 13
(U)SANS analyses for low porosity CM sample and low reactivity WD sample reacted for
2 days (a) and 45 days (b).
Figure 14 a-c shows the reacted rim on the dolostone in contact with NH4F for 4, 8 and 16 days. The rim of the reacted dolomite cross-sections is clearly visible to the naked eye but perhaps not as prominent as in the previously described limestone reactions. No secondary recrystallization rim is obvious to the naked eye. In addition, especially in the 8-day sample, the replacement reaction does not form a simple reaction front but moves inward via advantageous pathways. As the initial porosity and permeability of this rock are relatively low, however, it is not clear whether these are
50
ACS Paragon Plus Environment
Page 50 of 101
Page 51 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
controlled by pre-existing pathways within the rock, or represent wormhole-like, positivefeedback controlled pathways formed due to the new porosity caused by the reaction.73,74
Figure 14: Porosity changes during replacement of dolomite (WD) by fluorite as a function of reaction time. (a) 4-days reaction time, (b) 8-days and (c) 22-days reaction time, respectively. Porosity measurements were made using (U)SAXS.
Porosity analyses by (U)SAXS (Figure 14 a-c) did not show consistent porosity changes in the rim within the detectable pore size range. However, (U)SANS analyses showed a clear increase of the porosity with replacement (Figure 13), but this does not, 51
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
necessarily, correlate well with the optically observable replacement rim. After two days of replacement reaction, porosity increases near-linearly from approximately 7.1% in the core to approximately 18.5% at the rim, all of which are significantly greater than the 2.2% porosity observed in the starting material. Thus, despite the presence of the white rim near the edge of the core, it is clear that the reaction has proceeded to some extent throughout the entire core. The differences between the (U)SANS and (U)SAXS may be due to the averaging effects of the (U)SANS masks relative to the patchy replacement observed optically, whereas the highly localized (U)SAXS analysis may singly have only observed areas of no porosity change. From the optical observations, it is evident that the replacement proceeds along grain boundaries or other advantageous pathways. Possibly, the reaction proceeded throughout the core along these pathways, which are too small to be visible by the naked eye. Collected WAXS diffraction patterns of the replaced dolostone are more complicated to interpret in comparison to the limestone replacement characterization as two products (fluorite and sellaite) are expected to form. In nature, incomplete replacement of dolomite by fluorite was reported.75 52
ACS Paragon Plus Environment
Page 52 of 101
Page 53 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
As noted above, optically the WD samples show a somewhat patchy outer rim with an unreacted-looking inner core. The appearance of this rim is consistent with SEM-EDX observations and ToF-SIMS mapping (Figure 15), but the structural details of the reaction rim are quite distinct from the ones observed in the limestone replacement. The previously observed reaction rims were present in the form of a sharp boundary between the replaced and the original phase. Three different regions were identified in the dolomite samples reacted for 45 days (Figure 15). One is the original dolomite limestone as identified by chemical mapping. Then, there are two other sections of the reaction rim that are apparent (regions 2 and 3). At the interface between the parent and product phases, region 2, the material exhibits a higher BSE contrast compared to regions 1 and 3, indicating a different chemical composition. Its chemical composition is Ca, F, Mg in a homogeneous distribution with grains of high Si/K concentrations (see Supporting Information, Figure S18 for elemental maps). Mineralogically, this most likely corresponds to a solid solution of fluorite (CaF2) and sellaite (MgF2) due to the homogeneous distribution of both calcium, magnesium and fluorite in these regions. Furthermore, grains 53
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
(50-80 µm size) of a mineral phase rich in K, Al, Si and O (possibly orthoclase feldspars KAlSi3O8) were identified by SEM-EDS mapping (Supporting Information, Figures S18 and S19).
54
ACS Paragon Plus Environment
Page 54 of 101
Page 55 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
55
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Figure 15
SEM-BSE and ToF-SIMS imaging of dolomite samples reacted with NH4F for 45 days. (a)
BSE image showing three different regions consisting of original dolomite (1), partially replaced dolomite by fluoride (2) and brittle rim zone (3). Chemical mapping by ToF-SIMS (b + c) near the outside edge (edge visible near right lower corner) indicated that the replaced region contains fluoride and that replacement propagates faster along grain boundaries. Grains visible in the F-rich parts are Si-rich grains (see Supplementary information, Figure S20 for detailed elemental maps).
The outermost rim, region 3, which has approximately the same gray value as in BSE-SEM images as the pristine core, is highly fractured parallel to the core surface, and contains large pores. The chemo-mechanical fracturing visible at the edge of the rock core (Figure 15 a, region marked with 3) is the likely origin of the missing material visible in the 8 and 22-day experiments shown in Figure 14. As was the case with the CM samples, ToF-SIMS maps of the fluorine distribution in WD reacted for 45 days indicate that fluorine migrates faster via the grain boundaries. In the case of the dolomite replacement, additional pathways within the calcite crystals via twin boundaries were preferential replacement routes (Figure 16). It is evident in SEM-EDS mappings, that the fluoride is incorporated along grain boundaries, but also along twin boundaries, which are common in calcite crystals.76 56
ACS Paragon Plus Environment
Page 56 of 101
Page 57 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Figure 16
SEM-EDX mapping of the outer rim region (region 3 in Figure 15) of dolomite reacted with
NH4F for 45 days (WDF45). An outermost rim of homogeneous Ca-F-Mg-phases is present, whereas towards the middle the dolomite is replaced along grain boundaries and possibly twin boundaries within the grains.
3.3.
Comparison between the Effect of Chemical Reactivity and Microstructure on
Replacement Extent
To compare the effect of the microstructure and the chemical reactivity on the replacement extent and replacement rate, we used rim thicknesses determined via image analyses for CM and WD samples and calculated the replacement reaction rate based on reaction time. Based on WAXS patterns for CM limestone and on SEM-EDS mapping for WD dolostone, we can safely assume that the reaction rim visible in optical light is 57
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
replaced material. Therefore, we have used photographs of the thin section (see Supporting Information Figure S21) and determined the rim thickness by averaging four measurements at each of the sides of the rock core thin section (left, top, right and bottom of circular sample). For high porosity TC limestone, a complete reaction of the core was observed in WAXS, therefore the reaction length was used as the radius of the sample. It is important to note that this assumption will only provide a minimum replacement reaction rate. The individual measurement values are presented in Table 4.
Table 4
Measured reaction rim width (see Supporting Information, Figure S21 for images used for
image analyses).
Sample name CMF4 CMF8 CMF22 TCF4 TCF8 TCF22 WDF4 WDF8 WDF22
Reaction rim width right [cm] 0.224 0.338 0.320
0.092 0.100 0.146
Reaction rim width Reaction rim Reaction rim width - Reaction rim width - left [cm] width - top [cm] bottom [cm] average [cm] 0.261 0.206 0.202 0.223 0.356 0.360 0.349 0.351 0.364 0.348 0.334 0.342 1.588 complete replacement based on WAXS 1.588 1.588 0.144 0.127 0.073 0.109 0.089 0.164 0.077 0.108 0.110 0.118 0.124 0.125
58
ACS Paragon Plus Environment
Page 58 of 101
Page 59 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Based on the measured reaction rim widths, replacement reaction rates were calculated after the following scheme: First, the volume of the core was calculated using 𝑉𝑐𝑜𝑟𝑒 = 𝜋 𝑟𝑐𝑜𝑟𝑒2 × ℎ with 𝑟𝑐𝑜𝑟𝑒 = 0.79375 cm2. We set the core height h equal to one (h = 1 cm) as all of the cores have the same height.
To calculate the reaction rim volume (𝑉𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑟𝑖𝑚), we first calculated the inner, unreacted core volume (𝑉𝑢𝑛𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑐𝑜𝑟𝑒) using the equation (4):
𝑉𝑢𝑛𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑐𝑜𝑟𝑒 = 𝜋𝑟𝑢𝑛𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑐𝑜𝑟𝑒2 × ℎ
(4)
The radius for the unreacted core (𝑟𝑢𝑛𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑐𝑜𝑟𝑒) was calculated by subtracting the reacted rim width determined via image analyses (Table 4) from the core radius. The sample height was treated the same as for the core volume calculation, namely h = 1 cm. 𝑉𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑟𝑖𝑚 was calculated using equation (5):
𝑉𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑟𝑖𝑚 = 𝑉𝑐𝑜𝑟𝑒 ― 𝑉𝑢𝑛𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑐𝑜𝑟𝑒
59
ACS Paragon Plus Environment
(5)
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Page 60 of 101
The replacement rate 𝑅𝑟𝑒𝑝𝑙𝑎𝑐𝑒𝑚𝑒𝑛𝑡 was then calculated by dividing the volume of the reacted rim (𝑉𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑟𝑖𝑚) by the reaction time (𝑡𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛) converted to seconds (equation 6):
𝑅𝑟𝑒𝑝𝑙𝑎𝑐𝑒𝑚𝑒𝑛𝑡 = 𝑉𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑟𝑖𝑚 𝑡𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛
Table 5 Detailed results of replacement rate calculation.
Reaction rim Vcore runreacted core Vunreacted core Vreacted rim Sample name width - average 3 3 3 [cm] [cm ] [cm ] [cm ] [cm] CMF4 0.22325 0.571 1.022 0.957 CMF8 0.35075 0.443 0.617 1.363 CMF22 0.3415 0.452 0.643 1.337 WDF4 0.109 0.685 1.473 0.506 WDF8 0.1075 1.97933 0.686 1.479 0.500 WDF22 0.1245 0.669 1.407 0.572 TCF4 1.5875 TCF8 1.5875 not applicable as whole core reacted TCF22 1.5875
treaction [days]
treaction [s] 4 8 22 4 8 22
Rreplacement 3
[cm /s]
345600 691200 1900800 345600 691200 1900800
2.77E-06 1.97E-06 7.03E-07 1.46E-06 7.23E-07 3.01E-07
4 345600
5.727E-06
When comparing the relationship between replacement rate and starting porosity for the different materials (Figure 17), it is evident that a 4.6 times higher porosity leads to a change in replacement rate of half an order in magnitude. The reduced chemical reactivity does not change the replacement rate significantly. From the calculated rates for low 60
ACS Paragon Plus Environment
Page 61 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
porosity CM limestone and low porosity/low reactivity dolostone, it is evident that the reaction rate is not the same for all rock cores. This is most likely due to local inhomogeneities or higher availability of advantageous pathways.
Figure 17
Replacement reaction rate [cm3/s] as a dependence of the starting porosity for low/high
porosity limestone and dolostone.
61
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
4. Conclusion In conclusion, this study presents the comparison of the effect of porosity changes on replacement reactions with changes in the chemical reactivity and their effects. It is evident that a higher starting porosity accelerates the replacement reaction, as expected due to higher reactive surface area and more accessible pathways for the fluid to migrate into the rock. In the case of low porosity limestone replacement, grain boundaries are pathways into the rock core. During the replacement reactions, these grain boundary regions are dissolved first, and material is then reprecipitated in the shape of acicular crystals due to space constraints. In a second, textural reequlibration step, porosity is closed, and the acicular crystals are reformed into a more energy-efficient morphology. Furthermore, we were able to identify that changes in the high porosity limestone were mainly in the macropore regime, whereas the lower porosity limestone exhibits mainly changes on the nano-scale. Similar to the acicular fluorite growth, we hypothesize here that space constraints result in the formation of smaller pore sizes in the less porous limestone. 62
ACS Paragon Plus Environment
Page 62 of 101
Page 63 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
In the case of dolostone replacement, which we used as an example of lowered chemical reactivity, the reaction proceeded even slower and was more complex to interpret. Similar to the limestone, a reaction rim formed, however, this reaction rim was not completely matched to the porosity changes. By chemical composition mappings, we were able to identify two different regions of the reaction rim: a fluorine-rich region near the interface to the original rock, which consists most likely of a solid solution between fluorite (CaF2) and sellaite (MgF2). The outermost edge of the reacted rim exhibited higher fluorine concentrations along grain boundaries and twin boundaries. Due to the lower chemical reactivity, the replacement reaction was decelerated, offering an opportunity to capture these replacement pathways. Lastly, the preferential pathways available in a rock’s texture consist of fractures, connected pores, unconnected pores, grain boundaries and twin boundaries. Depending on the availability of these features, replacement reactions proceed along these advantageous pathways. Our results emphasize the importance of detailed microstructural and textural composition analyses when addressing replacement rates and reactions in experimental systems and in nature. 63
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Page 64 of 101
5. Acknowledgements This work was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and Biosciences Division. This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne
National
Laboratory
under
Contract
No.
DE-AC02-06CH11357.
We
acknowledge the support of the National Institute of Standards and Technology, Center for Neutron Research, U.S. Department of Commerce in providing the research neutron facilities used in this work. Access to both NBG30 SANS and BT5 USANS was provided by the Center for High Resolution Neutron Scattering, a partnership between the National Institute of Standards and Technology and the National Science Foundation under Agreement No. DMR-1508249. Certain commercial equipment, instruments, materials and software are identified in this paper to foster understanding. Such identification does not imply recommendation or endorsement by the National Institute of Standards and 64
ACS Paragon Plus Environment
Page 65 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Technology or the Department of Energy, nor does it imply that the materials or equipment identified are necessarily the best available for the purpose. A portion of this research used resources at the High Flux Isotope Reactor and Spallation Neutron Source, DOE Office of Science User Facilities operated by the Oak Ridge National Laboratory. We would like to thank M. Pearce and two anonymous reviewers for their constructive and detailed comments as well as editor S. Chakraborty for editorial comments.
65
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
AUTHOR INFORMATION Corresponding Author * Phone: ++1 865-576-7184; e-mail:
[email protected],
[email protected] Supplementary Information for Controls of Microstructure and Chemical Reactivity on the Replacement of Limestone by Fluorite Studied Using Spatially Resolved Small Angle Xray and Neutron Scattering: Porosity determination in reference materials, Pore Size distribution analysis in reference materials, wide-angle scattering spectra, SEM investigation of needle-shaped fluorite, detailed (U)SAXS and (U)SANS results.
6. References (1)
Altree-Williams, A.; Pring, A.; Ngothai, Y.; Brugger, J. Textural and Compositional Complexities Resulting from Coupled Dissolution–Reprecipitation Reactions in Geomaterials. Earth-Science Rev. 2015, 150, 628–651.
(2)
Putnis, A. Mineral Replacement Reactions. Rev. Mineral. Geochemistry 2009, 70 66
ACS Paragon Plus Environment
Page 66 of 101
Page 67 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
(1), 87–124.
(3)
Putnis, A. Mineral Replacement Reactions: From Macroscopic Observations to Microscopic
Mechanisms.
Mineral.
Mag.
2002,
66
(5),
689–708.
https://doi.org/10.1180/0026461026650056.
(4)
Borg, R. J.; Dienes, G. J. An Introduction to Solid State Diffusion; Elsevier, 2012.
(5)
Zhang, Y. Geochemical Kinetics; Princeton University Press, 2008.
(6)
Ruiz-Agudo, E.; Putnis, C. V; Putnis, A. Coupled Dissolution and Precipitation at Mineral–Fluid Interfaces. Chem. Geol. 2014, 383, 132–146.
(7)
Chakraborty, S.; Ganguly, J. Compositional Zoning and Cation Diffusion in Garnets. In Diffusion, Atomic Ordering, and Mass Transport; Springer, 1991; pp 120–175.
(8)
Fayek, M.; Anovitz, L. M.; Cole, D. R.; Bostick, D. A. O and H Diffusion in Uraninite: Implications for Fluid–Uraninite Interactions, Nuclear Waste Disposal, and Nuclear Forensics. Geochim. Cosmochim. Acta 2011, 75 (13), 3677–3686.
(9)
Chakraborty, S. Diffusion Coefficients in Olivine, Wadsleyite and Ringwoodite. Rev. 67
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Mineral. geochemistry 2010, 72 (1), 603–639. (10) Buening, D. K.; Buseck, P. R. Fe‐Mg Lattice Diffusion in Olivine. J. Geophys. Res. 1973, 78 (29), 6852–6862.
(11) Anovitz, L. M.; Elam, J. M.; Riciputi, L. R.; Cole, D. R. Isothermal Time‐series Determination of the Rate of Diffusion of Water in Pachuca Obsidian. Archaeometry 2004, 46 (2), 301–326.
(12) Anovitz, L. M.; Elam, J. M.; Riciputi, L. R.; Cole, D. R. The Failure of Obsidian Hydration Dating: Sources, Implications, and New Directions. J. Archaeol. Sci. 1999, 26 (7), 735–752.
(13) Emmanuel, S. Mechanisms Influencing Micron and Nanometer-Scale Reaction Rate Patterns during Dolostone Dissolution. Chem. Geol. 2014, 363, 262–269.
(14) Levenson, Y.; Emmanuel, S. Quantifying Micron-Scale Grain Detachment during Weathering Experiments on Limestone. Geochim. Cosmochim. Acta 2016, 173, 86–96.
68
ACS Paragon Plus Environment
Page 68 of 101
Page 69 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
(15) Israeli, Y.; Emmanuel, S. Impact of Grain Size and Rock Composition on Simulated Rock Weathering. Earth Surf. Dyn. 2018, 6 (2), 319–327.
(16) Levenson, Y.; Ryb, U.; Emmanuel, S. Comparison of Field and Laboratory Weathering Rates in Carbonate Rocks from an Eastern Mediterranean Drainage Basin. Earth Planet. Sci. Lett. 2017, 465, 176–183.
(17) Emmanuel, S.; Anovitz, L. M.; Day-Stirrat, R. J. Effects of Coupled ChemoMechanical Processes on the Evolution of Pore-Size Distributions in Geological Media. Rev. Mineral. Geochemistry 2015, 80 (1), 45–60.
(18) Hellmann, R.; Cotte, S.; Cadel, E.; Malladi, S.; Karlsson, L. S.; Lozano-Perez, S.; Cabié, M.; Seyeux, A. Nanometre-Scale Evidence for Interfacial DissolutionReprecipitation Control of Silicate Glass Corrosion. Nat. Mater. 2015, 14 (3), 307– 311. https://doi.org/10.1038/nmat4172.
(19) Geisler, T.; Nagel, T.; Kilburn, M. R.; Janssen, A.; Icenhower, J. P.; Fonseca, R. O. C.; Grange, M.; Nemchin, A. A. The Mechanism of Borosilicate Glass Corrosion
69
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Revisited.
Geochim.
Cosmochim.
Acta
Page 70 of 101
2015,
158,
112–129.
https://doi.org/10.1016/j.gca.2015.02.039.
(20) Pollok, K.; Putnis, C. V; Putnis, A. Mineral Replacement Reactions in Solid SolutionAqueous Solution Systems: Volume Changes, Reactions Paths and End-Points Using the Example of Model Salt Systems. Am. J. Sci. 2011, 311 (3), 211–236.
(21) Røyne, A.; Jamtveit, B. Pore-Scale Controls on Reaction-Driven Fracturing. Rev.
Mineral. Geochemistry 2015, 80, 25–44. https://doi.org/10.2138/rmg.2015.80.02. (22) Putnis, A. Transient Porosity Resulting from Fluid–Mineral Interaction and Its Consequences. Rev. Mineral. Geochemistry 2015, 80 (1), 1–23.
(23) Putnis, C. V; Geisler, T.; Schmid-Beurmann, P.; Stephan, T.; Giampaolo, C. An Experimental Study of the Replacement of Leucite by Analcime. Am. Mineral. 2007,
92 (1), 19–26. (24) Xia, F.; Brugger, J.; Chen, G.; Ngothai, Y.; O’Neill, B.; Putnis, A.; Pring, A. Mechanism and Kinetics of Pseudomorphic Mineral Replacement Reactions: A
70
ACS Paragon Plus Environment
Page 71 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Case Study of the Replacement of Pentlandite by Violarite. Geochim. Cosmochim.
Acta 2009, 73 (7), 1945–1969. (25) Harlov, D. E.; Wirth, R.; Förster, H.-J. An Experimental Study of Dissolution– Reprecipitation in Fluorapatite: Fluid Infiltration and the Formation of Monazite.
Contrib. to Mineral. Petrol. 2005, 150 (3), 268–286. (26) Kasioptas, A.; Perdikouri, C.; Putnis, C. V; Putnis, A. Pseudomorphic Replacement of Single Calcium Carbonate Crystals by Polycrystalline Apatite. Mineral. Mag. 2008, 72 (1), 77–80.
(27) Offeddu, F. G.; Cama, J.; Soler, J. M.; Putnis, C. V. Direct Nanoscale Observations of the Coupled Dissolution of Calcite and Dolomite and the Precipitation of Gypsum.
Beilstein J. Nanotechnol. 2014, 5 (1), 1245–1253. (28) Klasa, J.; Ruiz-Agudo, E.; Wang, L. J.; Putnis, C. V.; Valsami-Jones, E.; Menneken, M.; Putnis, A. An Atomic Force Microscopy Study of the Dissolution of Calcite in the Presence of Phosphate Ions. Geochim. Cosmochim. Acta 2013, 117, 115–128.
71
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Page 72 of 101
https://doi.org/10.1016/j.gca.2013.03.025.
(29) Putnis,
C.
V.;
Tsukamoto,
K.;
Nishimura,
Y.
Direct
Observations
of
Pseudomorphism: Compositional and Textural Evolution at a Fluid-Solid Interface.
Am. Mineral. 2005, 90 (11–12), 1909–1912. https://doi.org/10.2138/am.2005.1990. (30) Jonas, L.; Müller, T.; Dohmen, R.; Immenhauser, A.; Putlitz, B. Hydrothermal Replacement of Biogenic and Abiogenic Aragonite by Mg-Carbonates – Relation between Textural Control on Effective Element Fluxes and Resulting Carbonate Phase.
Geochim.
Cosmochim.
Acta
2017,
196,
289–306.
https://doi.org/10.1016/j.gca.2016.09.034.
(31) Pedrosa, E. T.; Putnis, C. V.; Putnis, A. The Pseudomorphic Replacement of Marble by Apatite: The Role of Fluid Composition. Chem. Geol. 2016, 425, 1–11. https://doi.org/10.1016/j.chemgeo.2016.01.022.
(32) Spruzeniece, L.; Piazolo, S.; Daczko, N. R.; Kilburn, M. R.; Putnis, A. Symplectite Formation in the Presence of a Reactive Fluid: Insights from Hydrothermal
72
ACS Paragon Plus Environment
Page 73 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Experiments.
J.
Metamorph.
Geol.
2017,
35
(3),
281–299.
https://doi.org/10.1111/jmg.12231.
(33) Anovitz, L. M.; Cole, D. R. Characterization and Analysis of Porosity and Pore Structures. Rev. Mineral. Geochemistry 2015, 80 (1), 61–164.
(34) Kolesnikov, A. I.; Reiter, G. F.; Choudhury, N.; Prisk, T. R.; Mamontov, E.; Podlesnyak, A.; Ehlers, G.; Seel, A. G.; Wesolowski, D. J.; Anovitz, L. M. Supplementary Information For Quantum Tunneling of Water in Beryl: A New State of
the
Water
Molecule.
Phys.
Rev.
Lett.
2016,
116
(16),
167802.
https://doi.org/10.1103/PhysRevLett.116.167802.
(35) Knight, A. W.; Tigges, A. B.; Ilgen, A. G. Adsorption of Copper (II) on Mesoporous Silica: The Effect of Nano-Scale Confinement. Geochem. Trans. 2018, 19 (1), 13. https://doi.org/10.1186/s12932-018-0057-4.
(36) Yuan, K.; Lee, S. S.; De Andrade, V.; Sturchio, N. C.; Fenter, P. Replacement of Calcite (CaCO 3 ) by Cerussite (PbCO 3 ). Environ. Sci. Technol. 2016, 50 (23),
73
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Page 74 of 101
12984–12991. https://doi.org/10.1021/acs.est.6b03911.
(37) Yuan, K.; De Andrade, V.; Feng, Z.; Sturchio, N. C.; Lee, S. S.; Fenter, P. Pb 2+ Calcite
Interactions
under
Far-from-Equilibrium
Conditions:
Formation
of
Micropyramids and Pseudomorphic Growth of Cerussite. J. Phys. Chem. C 2018,
122 (4), 2238–2247. https://doi.org/10.1021/acs.jpcc.7b11682. (38) Yuan, K.; Lee, S. S.; Wang, J.; Sturchio, N. C.; Fenter, P. Templating Growth of a Pseudomorphic Lepidocrocite Microshell at the Calcite-Water Interface. Chem.
Mater. 2018, 30 (3), 700–707. https://doi.org/10.1021/acs.chemmater.7b03921. (39) Putnis, A.; Putnis, C. V. The Mechanism of Reequilibration of Solids in the Presence of a Fluid Phase. J. Solid State Chem. 2007, 180 (5), 1783–1786.
(40) Qian, G.; Xia, F.; Brugger, J.; Skinner, W. M.; Bei, J.; Chen, G.; Pring, A. Replacement of Pyrrhotite by Pyrite and Marcasite under Hydrothermal Conditions up to 220 C: An Experimental Study of Reaction Textures and Mechanisms. Am.
Mineral. 2011, 96 (11–12), 1878–1893.
74
ACS Paragon Plus Environment
Page 75 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
(41) Wu, W.; Ma, Y.; Xing, Y.; Zhang, Y.; Yang, H.; Luo, Q.; Wang, J.; Li, B.; Qi, L. CaDoped Strontianite–Calcite Hybrid Micropillar Arrays Formed via Oriented Dissolution and Heteroepitaxial Growth on Calcite. Cryst. Growth Des. 2015, 15 (5), 2156–2164.
(42) Jonas, L.; Müller, T.; Dohmen, R.; Baumgartner, L.; Putlitz, B. Transport-Controlled Hydrothermal Replacement of Calcite by Mg-Carbonates. Geology 2015, 43 (9), 779–783. https://doi.org/10.1130/G36934.1.
(43) Brown, W. L.; Parsons, I. Exsolution and Coarsening Mechanisms and Kinetics in an Ordered Cryptoperthite Series. Contrib. to Mineral. Petrol. 1984, 86 (1), 3–18.
(44) Kelemen, P. B.; Matter, J.; Streit, E. E.; Rudge, J. F.; Curry, W. B.; Blusztajn, J. Rates and Mechanisms of Mineral Carbonation in Peridotite: Natural Processes and Recipes for Enhanced, in Situ CO2 Capture and Storage. Annu. Rev. Earth
Planet. Sci. 2011, 39 (1), 545–576. https://doi.org/10.1146/annurev-earth-092010152509.
75
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
(45) Dobson, P. F.; Kneafsey, T. J.; Hulen, J.; Simmons, A. Porosity, Permeability, and Fluid Flow in the Yellowstone Geothermal System, Wyoming. J. Volcanol.
Geotherm. Res. 2003, 123 (3–4), 313–324. (46) Qian, G.; Brugger, J.; Skinner, W. M.; Chen, G.; Pring, A. An Experimental Study of the Mechanism of the Replacement of Magnetite by Pyrite up to 300 C. Geochim.
Cosmochim. Acta 2010, 74 (19), 5610–5630. (47) Pedrosa, E. T.; Putnis, C. V.; Renard, F.; Burgos-Cara, A.; Laurich, B.; Putnis, A. Porosity Generated during the Fluid-Mediated Replacement of Calcite by Fluorite.
CrystEngComm 2016, 18 (36), 6867–6874. https://doi.org/10.1039/c6ce01150k. (48) Pedrosa, E. T.; Boeck, L.; Putnis, C. V.; Putnis, A. The Replacement of a Carbonate Rock by Fluorite: Kinetics and Microstructure. American Mineralogist. 2017, pp 126–134. https://doi.org/10.2138/am-2017-5725.
(49) Ball, J. W.; Nordstrom, D. K. User’s Manual for WATEQ4F, with Revised Thermodynamic Data Base and Test Cases for Calculating Speciation of Major,
76
ACS Paragon Plus Environment
Page 76 of 101
Page 77 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Trace, and Redox Elements in Natural Waters. 1991.
(50) Davis, K. J.; Dove, P. M.; De Yoreo, J. J. The Role of Mg2+ as an Impurity in Calcite
Science
Growth.
(80-.
).
290
2000,
(5494),
1134–1137.
https://doi.org/10.1126/science.290.5494.1134.
(51) Astilleros, J. M.; Fernández-Díaz, L.; Putnis, A. The Role of Magnesium in the Growth of Calcite: An AFM Study. Chem. Geol. 2010, 271, 52–58.
(52) Anovitz, L. M.; Cole, D. R.; Fayek, M. Mechanisms of Rhyolitic Glass Hydration below the Glass Transition. Am. Mineral. 2008, 93 (7), 1166–1178.
(53) Anovitz, L. M.; Lynn, G. W.; Cole, D. R.; Rother, G.; Allard, L. F.; Hamilton, W. A.; Porcar, L.; Kim, M.-H. A New Approach to Quantification of Metamorphism Using Ultra-Small and Small Angle Neutron Scattering. Geochim. Cosmochim. Acta 2009,
73 (24), 7303–7324. (54) Ilavsky, J.; Zhang, F.; Andrews, R. N.; Kuzmenko, I.; Jemian, P. R.; Levine, L. E.; Allen,
A.
J.
Development
of
Combined
Microstructure
77
ACS Paragon Plus Environment
and
Structure
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Page 78 of 101
Characterization Facility for in Situ and Operando Studies at the Advanced Photon Source. J. Appl. Crystallogr. 2018, 51 (3), 867–882.
(55) Ilavsky, J.; Jemian, P. R. Irena: Tool Suite for Modeling and Analysis of Small-Angle Scattering. J. Appl. Crystallogr. 2009, 42 (2), 347–353.
(56) Anovitz, L. M.; Freiburg, J. T.; Wasbrough, M.; Mildner, D. F. R.; Littrell, K. C.; Pipich, V.; Ilavsky, J. The Effects of Burial Diagenesis on Multiscale Porosity in the St. Peter Sandstone: An Imaging, Small-Angle, and Ultra-Small-Angle Neutron Scattering
Analysis.
Mar.
Pet.
Geol.
2017.
https://doi.org/https://doi.org/10.1016/j.marpetgeo.2017.11.004.
(57) Wang, H.-W.; Anovitz, L. M.; Burg, A.; Cole, D. R.; Allard, L. F.; Jackson, A. J.; Stack, A. G.; Rother, G. Multi-Scale Characterization of Pore Evolution in a Combustion Metamorphic Complex, Hatrurim Basin, Israel: Combining (Ultra) Small-Angle Neutron Scattering and Image Analysis. Geochim. Cosmochim. Acta 2013, 121, 339–362. https://doi.org/10.1016/j.gca.2013.07.034.
78
ACS Paragon Plus Environment
Page 79 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
(58) Anovitz, L. M.; Cole, D. R.; Rother, G.; Valley, J. W.; Jackson, A. Analysis of NanoPorosity in the St. Peter Sandstone Using (Ultra) Small Angle Neutron Scattering.
Geochim. Cosmochim. Acta 2010, 74 (12), A26–A26. (59) Anovitz, L. M.; Cole, D. R.; Rother, G.; Allard, L. F.; Jackson, A. J.; Littrell, K. C. Diagenetic Changes in Macro-to Nano-Scale Porosity in the St. Peter Sandstone: An (Ultra) Small Angle Neutron Scattering and Backscattered Electron Imaging Analysis. Geochim. Cosmochim. Acta 2013, 102, 280–305.
(60) Anovitz, L. M.; Rother, G.; Cole, D. R. Characterization of Rock Pore Features in Geothermal Systems Using Small Angle Neutron Scattering (SANS). In Proc 36th
Workshop on Geothermal Res Eng Sgp-Tr-191; 2011; pp 571–582. (61) Radliński, A. P.; Boreham, C. J.; Lindner, P.; Randl, O.; Wignall, G. D.; Hinde, A.; Hope, J. M. Small Angle Neutron Scattering Signature of Oil Generation in Artificially and Naturally Matured Hydrocarbon Source Rocks. Org. Geochem. 2000, 31 (1), 1–14.
79
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
(62) Radlinski, A. P. Small-Angle Neutron Scattering and the Microstructure of Rocks.
Rev. Mineral. Geochemistry 2006, 63 (1), 363–397. (63) Barker, J. G.; Glinka, C. J.; Moyer, J. J.; Kim, M. H.; Drews, A. R.; Agamalian, M. Design and Performance of a Thermal-Neutron Double-Crystal Diffractometer for USANS at NIST. J. Appl. Crystallogr. 2005, 38 (6), 1004–1011.
(64) Agamalian, M.; Heroux, L.; Littrell, K. C.; Carpenter, J. M. Progress on The Timeof-Flight Ultra Small Angle Neutron Scattering Instrument at SNS. In Journal of
Physics: Conference Series; IOP Publishing, 2018; Vol. 1021, p 12033. (65) Kline, S. R. Reduction and Analysis of SANS and USANS Data Using IGOR Pro.
J. Appl. Crystallogr. 2006, 39 (6), 895–900. (66) Glinka, C. J.; Barker, J. G.; Hammouda, B.; Krueger, S.; Moyer, J. J.; Orts, W. J. The 30 m Small‐angle Neutron Scattering Instruments at the National Institute of Standards and Technology. J. Appl. Crystallogr. 1998, 31 (3), 430–445.
(67) Wignall, G. D.; Littrell, K. C.; Heller, W. T.; Melnichenko, Y. B.; Bailey, K. M.; Lynn,
80
ACS Paragon Plus Environment
Page 80 of 101
Page 81 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
G. W.; Myles, D. A.; Urban, V. S.; Buchanan, M. V; Selby, D. L. The 40 m General Purpose Small-Angle Neutron Scattering Instrument at Oak Ridge National Laboratory. J. Appl. Crystallogr. 2012, 45 (5), 990–998.
(68) Merritt, M.; Zhang, Y. Interior-Point Gradient Method for Large-Scale Totally Nonnegative Least Squares Problems. J. Optim. Theory Appl. 2005, 126 (1), 191– 202.
(69) Nelson, A. Co-Refinement of Multiple-Contrast Neutron/X-Ray Reflectivity Data Using MOTOFIT. J. Appl. Crystallogr. 2006, 39 (2), 273–276.
(70) Beaucage, G. Approximations Leading to a Unified Exponential/Power-Law Approach to Small-Angle Scattering. J. Appl. Crystallogr. 1995, 28 (6), 717–728.
(71) Pearce, M. A.; Timms, N. E.; Hough, R. M.; Cleverley, J. S. Reaction Mechanism for the Replacement of Calcite by Dolomite and Siderite: Implications for Geochemistry, Microstructure and Porosity Evolution during Hydrothermal Mineralisation.
Contrib.
to
Mineral.
Petrol.
81
ACS Paragon Plus Environment
2013,
166
(4),
995–1009.
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Page 82 of 101
https://doi.org/10.1007/s00410-013-0905-2.
(72) Etschmann, B.; Brugger, J.; Pearce, M. A.; Ta, C.; Brautigan, D.; Jung, M.; Pring, A. Grain Boundaries as Microreactors during Reactive Fluid Flow: Experimental Dolomitization of a Calcite Marble. Contrib. to Mineral. Petrol. 2014, 168 (2), 1–12.
(73) Ladd, A.; Starchenko, V.; Dutka, F.; Szymczak, P. Dissolution at the Pore Scale: Comparing Simulations and Experiments. In Bulletin of the American Physical
Society; APS, 2018; Vol. 63, p 9. (74) Starchenko, V.; Ladd, A. J. C. The Development of Wormholes in Laboratory-Scale Fractures: Perspectives From Three-Dimensional Simulations. Water Resour. Res. 2018, 54 (10), 7946–7959. https://doi.org/10.1029/2018WR022948.
(75) Zeeh, S. Complex Replacement of Saddle Dolomite by Fluorite within Zebra Dolomites - An Example from Radnig, Carinthia, Austria. Miner. Depos. 1995, 30 (6), 469–475. https://doi.org/10.1007/BF00196406.
(76) Bruno,
M.;
Massaro,
F.
R.;
Rubbo,
M.;
82
ACS Paragon Plus Environment
Prencipe,
M.;
Aquilano,
D.
Page 83 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
(10.4),(01.8),(01.2), and (00.1) Twin Laws of Calcite (CaCO3): Equilibrium Geometry of the Twin Boundary Interfaces and Twinning Energy. Cryst. Growth
Des. 2010, 10 (7), 3102–3109. Table of Contents Graphic
83
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Scanning Electron Microscopy Backscattered Electron (SEM-BSE) contrast images of the starting materials. (a) High porosity TC limestone, (b) low porosity CM limestone and (c) low porosity WD dolostone. The low porosity WD dolostone contains dolomite (dol), orthoclase (or) and quartz (qtz) as identified by XRD and SEM-EDS mappings (see Supporting Information, Figure S1 for mappings.) 214x466mm (150 x 150 DPI)
ACS Paragon Plus Environment
Page 84 of 101
Page 85 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Radial Cd masks designed for this experiment. 144x147mm (96 x 96 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
(U)SANS characterization after 2 and 45 days. (a) (U)SANS analyses of average porosity in distance from the center of the core for low porosity CM and high porosity TC limestone reacted for 2 days. Porosity increases at the rim to 14 % for both limestones. (b) (U)SANS analyses of the porosity in limestones reacted for 45 days showing a reduction of porosity in the outer rim of both high and low porosity limestone. 244x90mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 86 of 101
Page 87 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
SEM characterization of CM limestone replaced by fluorite after (a) 8, (b) 22 and (c) 45 days. Near the outer rim of the samples, a region of reduced porosity is visible (marked by 1). At the rim between limestone (Calc) and fluorite (Fl), acicular fluorite growth is visible (2). 118x268mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Porosity distribution determined by (U)SANS for (a) low porosity CM limestone and (b) high porosity TC limestone reacted for 2, 8, 22 and 45 days. 258x89mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 88 of 101
Page 89 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Porosity changes in CM limestone with increasing reaction time (a) 4 days reaction time, (b) 8 days reaction time and (c) 22 days reaction time. Diagrams show porosity [%], Mean and Median Diameter [µm] for each (U)SAXS measurement point collected during linescan traversing the middle of the core indicated in (a), (b) and (c). Color code for measurement points indicates phases identified by WAXS pattern (for individual WAXS pattern see Supporting Information, Figure S8 and S9 and Table S3, S4 and S5), with orange indicating calcite, red indicating fluorite and pale yellow indicating a calcite/fluorite mix. Note the recrystallized area near the rim after 22 days. 168x116mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Porosity changes in TC limestone with increasing reaction time (a) 4 days reaction time, (b) 8 days reaction time and (c) 22 days reaction time. Diagrams show porosity [%] on the left y-axis and on the right y-axis mean and median diameter [µm] for each (U)SAXS measurement point collected during linescan traversing the middle of the core. The orange color code indicates that the measurement points were identified in WAXS pattern as fluorite (see Supporting Information, Figure S13 and S14, Table S6, S7 and S8 for individual patterns). Again, note the recrystallization of the outer edge. 175x115mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 90 of 101
Page 91 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Pore Size Distribution analyzed by (U)SAXS for low porosity CM limestone replaced with fluorite for 4 days (a), 8 days (b) and 22 days (c). Sample regions probed by (U)SAXS which are fluorite as indicated by WAXS are depicted in red, calcite is green and a mixture of the two phases is blue. Sample locations are the same as in Figure 8. 668x899mm (150 x 150 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
BSE-SEM images of high porosity TC and low porosity CM limestone at day 0, 2 and 45 of replacement reaction. The high porosity TC limestone (a) shows a reaction rim after 2 days of replacement (c) and is completely replaced after 45 days (e). Higher porosity in the replaced region is visible. The low porosity CM limestone (b) shows a clear reaction front indicating enhanced replacement along the grain boundaries after 2 days (d) and after 45 days (f). 224x245mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 92 of 101
Page 93 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
: SEM and SEM-EDX characterization of the replacement for the low porosity CM limestone. (a) BSE image of low porosity CM limestone replaced for 45 days (CMF45) (b) Si distribution (c) Ca distribution and (d) F distribution. Enhanced F incorporation along the grain boundaries is visible. 231x181mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Two types of interface between original phase and product phase imaged SEM-BSE contrast images of low porosity limestone reacted for 2 days (CMF2). (a) Interface with a gap between parent and product phase (b) Acicular fluorite crystal growth along preferential pathways. 102x181mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 94 of 101
Page 95 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Chemical mapping of the interface between limestone and fluorite in TCF2 sample showing uniform incorporation of fluorine into limestone. The higher porosity of fluorite-replaced regions is visible in (a) with rectangular mapping area (b) marked. Mappings for Si (c), F (d) and Ca (e) show incorporation of fluorine into the limestone. 221x95mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
(U)SANS analyses for low porosity CM sample and low reactivity WD sample reacted for 2 days (a) and 45 days (b). 304x109mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 96 of 101
Page 97 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
Porosity changes during replacement of dolomite (WD) by fluorite as a function of reaction time. (a) 4-days reaction time, (b) 8-days and (c) 22-days reaction time, respectively. According porosity measurements using (U)SAXS. 232x154mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
SEM-BSE and ToF-SIMS imaging of dolomite samples reacted with NH4F for 45 days. (a) BSE image showing three different regions consisting of original dolomite (1), partially replaced dolomite by fluoride (2) and brittle rim zone (3). Chemical mapping by ToF-SIMS (b + c) near the outside edge (edge visible near right lower corner) indicated that the replaced region contains fluoride and that replacement propagates faster along grain boundaries. Grains visible in the F-rich parts are Si-rich grains (see Supplementary information, Figure S20 and S21 for detailed elemental maps). 139x177mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 98 of 101
Page 99 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
SEM-EDX mapping of the outer rim region (region 3 in Figure 15) of dolomite reacted with NH4F for 45 days (WDF45). An outermost rim of homogeneous Ca-F-Mg-phases is present, whereas towards the middle the dolomite is replaced along grain boundaries and possibly twin boundaries within the grains. 201x158mm (300 x 300 DPI)
ACS Paragon Plus Environment
ACS Earth and Space Chemistry 1 2 3 4 5 6 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
Replacement reaction rate [cm3/s] as a dependence of the starting porosity for low/high porosity limestone and dolostone. 232x197mm (150 x 150 DPI)
ACS Paragon Plus Environment
Page 100 of 101
Page 101 of 101 1 2 3 4 5 6 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
ACS Earth and Space Chemistry
TOC 254x146mm (300 x 300 DPI)
ACS Paragon Plus Environment