Ribose onto Brucite - ACS Publications

9. 3400 N Charles St, Baltimore MD, 21218 USA. 10. †. Present Address: School of Earth and Space Exploration, Arizona State University, 781 E Terrac...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Article

Cooperative and Inhibited Adsorption of DRibose onto Brucite [Mg(OH)] with Divalent Cations 2

Charlene Fae Estrada, Alyssa K Adcock, Dimitri A. Sverjensky, and Robert M Hazen ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.7b00095 • Publication Date (Web): 24 Oct 2017 Downloaded from http://pubs.acs.org on October 28, 2017

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

ACS Earth and Space Chemistry 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 33

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

To be submitted to ACS Earth and Space Chemistry

2

Cooperative and Inhibited Adsorption of D-

3

Ribose onto Brucite [Mg(OH)2] with Divalent

4

Cations

5 6

Charlene F. Estrada*,1,2,†, Alyssa K. Adcock3,⊥, Dimitri A. Sverjensky1

7

and Robert M. Hazen2

8 1

9

Johns Hopkins University, Department of Earth & Planetary Sciences,

10 11

3400 N Charles St, Baltimore MD, 21218 USA †

Present Address: School of Earth and Space Exploration, Arizona State University, 781 E Terrace Mall, Tempe,

12

AZ 85287, USA. 2

13 14

5251 Broad Branch Rd NW, Washington DC 20015 USA 3

15 16 17

Geophysical Laboratory, Carnegie Institution for Science,

⊥Present

Jacobs University, Campus Ring 1, 28759 Bremen, Germany

Address: Department of Chemistry, Georgetown University, 3700 O St NW, Washington DC 20057, USA *Corresponding Author email address: [email protected] (C. Estrada).

18

ACS Paragon Plus Environment

1

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 2 of 33

19

Abstract

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

The adsorption and concentration of sugars onto mineral surfaces in geochemical environments, such as hydrothermal systems, may have influenced the evolution of early life on Earth. We conducted batch adsorption experiments between D-ribose and brucite [Mg(OH)2], a mineral produced from serpentinite-hosted hydrothermal systems, over variable initial ribose concentrations at four ionic strengths resulting from different Mg2+ and Ca2+ ion concentrations in the aqueous phase. Ribose adsorption generally increased with greater initial concentration, and up to 0.3 µmol•m-2 ribose attached onto brucite with 0.6 mM Mg2+ present. Ribose adsorption decreased over six-fold (4.9x10-2 µmol•m-2) when the total Mg2+ ion concentration increased to 5.8 mM. Ribose adsorption increased to 0.4 µmol•m-2 when 4.2 mM CaCl2 was added to the system. Substantial amounts (over 21 µmol•m-2) of dissolved Ca also attached to the brucite surface independent of ribose concentration. We characterized the interactions between ribose, Ca, and the brucite surface by fitting a surface complexation model to adsorption data. We propose three types of surface reactions that were consistent with the experimental data and involve 1) a bidentate outer-sphere or a “standing” ribose surface species, 2) a monodentate Ca-ribose outer-sphere species, and 3) a monodentate Ca outer-sphere species. Our model predicts brucite particle surface charge is negative at low Mg2+ concentrations and further decreases upon the addition of MgCl2, which may hinder our proposed surface complexation of the ribose species, Rib-. We predict that brucite becomes positivelycharged with CaCl2 addition, which may be a consequence of the significant extent of Ca adsorption. The increase in ribose adsorption with CaCl2 is likely driven by Ca attachment and the formation of a positively-charged, cooperative Ca-ribose species that our model predicts will predominate over the “standing” ribose species on brucite. Our model of the ribose-brucite system, established by a combination of batch adsorption experiments and surface complexation modeling, has enabled predictions of ribose adsorption over a wide range of pH and complex environmental conditions.

47

Keywords: Ribose; Brucite; Surface Complexation; Mineral Surface Chemistry; Calcium;

48

RNA Evolution

49 50

1. Introduction

51

The interactions between organic molecules and the mineral-water interface may further

52

our current understanding of fundamental geological, biological, and atmospheric processes. For

53

instance, the reaction of organic matter, water, and minerals at the soil profile significantly

54

affects chemical weathering, the global carbon cycle, and atmospheric concentrations of CO21,2.

ACS Paragon Plus Environment

2

Page 3 of 33

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

Complex suites of biomolecules have condensed from inorganic precursors among gas, aqueous

56

fluids, and a limited assemblage of minerals within carbonaceous chondrites traveling through

57

space3. Microbes are closely associated with minerals of a wide variety of crystallographic habits

58

following biomineralization4-6. Filamentous microfossils cemented by iron oxides and silica have

59

been preserved at timescales exceeding one billion years7. Biomolecule interactions at mineral

60

surfaces have also been proposed to contribute to hypotheses on the origins of life on early Earth,

61

in which prebiotic molecules may have concentrated at the mineral-water interface prior to

62

assembling into more biochemically complex macromolecules such as RNA, DNA, and

63

proteins8,9. The emergence of RNA and DNA from nucleic acids and sugars is a fundamental

64

process in such an origins of life scenario8,10,11. Clay surfaces, such as montmorillonite, catalyze

65

the formation of ribonucleotide oligomer chains11,12. The mineral-water interface is a potentially

66

important local environment in the formation of more complex biomolecules because the

67

adsorbing organic species may typically form in water under the pH conditions or salinity of

68

interest11-13. Thus, the mineral-water interface may increase the availability of a particular

69

organic species that could contain more polarized or charged functional groups, and the

70

concentration of such groups may facilitate reactions among other sorbed biomolecules.

71

The monosaccharide

D-ribose

(C5H10O5) forms the backbone of RNA and DNA

72

molecules, and in water it forms an equilibrium mixture of five neutral isomers, among which

73

the β-ribopyranose predominates14. Researchers have previously synthesized ribose and other

74

sugars under plausible early Earth conditions15-17. LaRowe and Regnier18 calculated that abiotic

75

ribose formation under elevated pressures and temperatures up to 150 ºC and a bulk water

76

composition similar to that of ultramafic hydrothermal systems19,20 is thermodynamically

ACS Paragon Plus Environment

3

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 4 of 33

77

favorable18. Such a determination may hold implications for the evolution of more complex

78

organic compounds that contain ribose within hydrothermal environments.

79

A number of studies have focused on the adsorption of ribonucleotides onto mineral

80

surfaces, such as nontronite21, montmorillonite11,12,21-23, kaolinite23,24, α-Al2O325, SiO226,

81

Al(OH)324,27, manganese oxides28, allophane29, and TiO230. In the latter study, ribonucleotides

82

attached onto TiO2 to a greater extent compared with deoxyribonucelotides, and the authors

83

suggested that the hydroxyl groups of the ribose moiety might have enhanced ribonucleotide

84

adsorption at the mineral-water interface. It is possible that this observation holds implications

85

for the selection of nucleic acids with a ribose base. However, in comparison to investigations on

86

ribonucleic acids, ribose attachment at mineral surfaces has received relatively little attention,

87

presumably due to the low (< 5 %) extent of adsorption that is difficult to characterize as a

88

function of different environmental conditions. A notable exception is the work of Hashizume

89

and coworkers23,29, which reported ribose attachment onto the surfaces of montmorillonite and

90

allophane using a total organic carbon detector in the ppb range. However, in both studies, very

91

little (10

132

mM) of Ca or organic compounds might affect the dissolution of brucite in aqueous solution, and

133

accordingly, we studied relatively low (SOH2+>SOH_Rib-

(4b)

243

resulting in the overall reaction

244

2>SOH + HRib = >SOH2+>SOH_Rib–

245

and two surface sites are protonated in the second reaction:

246

HRib = Rib- + H+

(5a)

247

2>SOH + Rib- + 2H+ = 2(>SOH2+)_Rib-

(5b)

248

resulting in the overall reaction

249

2>SOH + H+ + HRib = 2(>SOH2+)_Rib–

250

in which “>SOH” represents a single brucite site at the solid-water interface (see Supporting

251

Information). In Figure 1a, both of these reactions resulted in close fits to the adsorption data

252

within experimental uncertainty, which provided support for the use of either eq 4 or 5 to

253

describe ribose adsorption onto brucite. The two surface reactions might result in a bidentate

254

outer-sphere complex, or a “standing” complex, at the brucite surface. Assuming that an outer-

255

sphere complex requires hydrogen bonding between a hydroxyl functional group of ribose and

256

the brucite surface, we illustrated a schematic representation of both surface complexes in Figure

257

2a,b.

258 259

(4c)

(5c)

If we express the neutral ribose molecule as H2Rib, we can rewrite eq 4 as H2Rib = Rib- + 2H+

(6a)

ACS Paragon Plus Environment

11

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 12 of 33

260

2>SOH + 2H+ + Rib- = 2(>SOH2+)_Rib-

(6b)

261

2(>SOH2+)_Rib- = >SOH2+>SRib- + H2O

(6c)

262

resulting in the overall reaction

263

2>SOH + H2Rib = >SOH2+>SRib– + H2O

264

in which the ribose molecule might adsorb as a bidentate complex with one inner-sphere and one

265

outer-sphere attachment. The ETLM cannot distinguish between these two surface reactions.

266

However, inner-sphere complexation might be less likely given the relatively short time for

267

ribose to attain a steady state and its relatively small amount of adsorption onto brucite in

268

comparison to aspartate, which we have predicted to form an inner-sphere complex on brucite35.

269 270 271 272

(6d)

The reactions in eqs 4 and 5 correspond with the equilibrium constants ∗ $ K _ !"# 

=

%&' %&'_()*# 

 %&' ()*

-./0,4

10 .4 4(5

(7)

and ∗ $ K 6()_ !"# 

(%&' )_()*#

= 



%&' ()* 

-./0,5

10 .4 4(5

(8)

273

The superscript “*” indicates that these reactios are expressed relative to the neutral >SOH

274

surface and “0” refers to a hypothetical 1.0 molal standard state55. The values ∆ψr,4 and ∆ψr,5

275

represent the electrical work involved in eqs 4 and 5, respectively. The electrical work includes a

276

contribution from the movement of water dipoles off the brucite surface according to ∆ψr = -

277

nH2O(ψ0- ψβ), where nH2O is the number of water molecules on the right-hand side of the reaction.

278

In eqs 4-5, nH2O = 0, which results in ∆ψr,4 = ψ0-ψβ and ∆ψr,5 = 2ψ0 - ψβ.

279

3.2. The Low-Ca2+ and High-Ca2+ Experiments

280

Ribose adsorption isotherms for the low-Ca2+ and high-Ca2+ experiments are shown in

281

Figure 1b. Γads was between 1.0x10-2 and 0.3 µmol•m-2 for the low-Ca2+ experiment. This

282

amount generally corresponded with 4.4 to 2.5 % ribose adsorption and generally decreased with

ACS Paragon Plus Environment

12

Page 13 of 33

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

283

increasing Γads. Ribose adsorption was significantly greater in the high-Ca2+ compared with the

284

Mg2+ experiments. Γads was between 4.8x10-2 and 0.4 µmol•m-2, and this range generally

285

corresponded with 11.4 to 3.8 % ribose adsorption. We also observed substantial amounts of Ca

286

adsorption onto brucite independent of ribose concentration (Figure 3). The amount of Ca

287

adsorption averaged 21.3 ± 1.3 µmol•m-2 (39.9 ± 2.8 %) for the low-Ca2+ experiment and 22.7 ±

288

3.5 µmol•m-2 (10.4 ± 1.6 %) for the high-Ca2+ experiment. All Ca adsorption data is reported in

289

Table S2.

290

We proposed that two surface reactions in combination with the “standing” ribose

291

complex represented by eqs 4 and 5 characterized the adsorption of Ca and ribose on brucite.

292

The first surface reaction involves the formation of a Ca-ribose surface complex:

293

HRib + Ca2+ = CaRib+ + H+

(9a)

294

>SOH + CaRib+ = >SOH_CaRib+

(9b)

295

resulting in the overall reaction

296

>SOH + HRib + Ca2+ = >SOH_CaRib+ + H+

297

We can interpret this reaction to involve the adsorption of a monodentate outer-sphere species on

298

the brucite surface (Figure 4a). The second reaction involves the adsorption of a hydrated Ca

299

complex:

300

Ca2+ + 2H2O = Ca(OH)2 + 2H+

(10a)

301

>SOH + H+ + Ca(OH)2 = >SOH2+_Ca(OH)2

(10b)

302

resulting in the overall reaction

303

>SOH + Ca2+ + 2H2O = >SOH2+_ Ca(OH)2 + H+

304

This reaction also involves a monodentate outer-sphere surface species shown in Figure 4b.

(9c)

(10c)

ACS Paragon Plus Environment

13

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 14 of 33

305

In Figure 4a, we speculated that the Ca-ribose outer-sphere species might attach through

306

a hydrogen bond between a deprotonated ribose hydroxyl group and the neutral brucite surface.

307

The Ca2+ ion might attach to 3 oxygen atoms of a single α-furanose ribose molecule, as

308

suggested by previous studies of Ca-ribose aqueous complexes51,56. The equilibrium constant of

309

the surface reaction in eq 9 can be expressed as

310 311 312 313

∗ $ K _( !")

=

%&'_:;(()*)  %&' ()* :;

-./0,9

10 .4 4(5

(10)

where ∆ψr,9 = +ψβ. We expressed the equilibrium constant of the surface reaction proposed in eq 10 as ∗ $ K _() 

=

%&' _:;(')   %&' :;  ' 

-./0,10

10 .4 4(5

(11)

314

where ∆ψr,10 = +ψ0. We proposed the reaction in eq 10 in a previous study of the brucite-

315

aspartate system35. Because the same batch of brucite was used in this study, we have adopted

316

the same equilibrium constant recalculated for the decreased site density of Ns = 31 sites•nm-2

317

(see Table 2).

318

We combined the surface reactions in eqs 4 and 5 and eqs 9 and 10 to fit the ribose

319

adsorption data obtained from the low-Ca2+ and high-Ca2+ experiments in Figure 1b. The

320

calculated curves are consistent with all the adsorption data collected from both Ca2+

321

experiments within analytical uncertainty. The proposed surface reactions also predict Ca

322

adsorption percentage data within error, as demonstrated in Figure 3. Therefore, we considered

323

the models predicted by the surface reactions in eqs 4 and 5 and eqs 9 and 10 to sufficiently

324

characterize ribose and Ca adsorption at the brucite-water interface at different ionic strengths,

325

pH conditions, and ribose concentrations.

326

3.4. Prediction of Brucite Particle Surface Charge

ACS Paragon Plus Environment

14

Page 15 of 33

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

327

The decrease in ribose surface adsorption in the high-Mg2+ experiment and its increased

328

adsorption in the high-Ca2+ experiment might be related to the electrostatic environment at the

329

brucite surface. We used the ETLM to predict brucite particle surface charge (ζ-potential, in mV),

330

which we assumed was equal to ΨD, otherwise known as the potential beginning at the diffuse

331

layer of the brucite-water interface. This prediction is shown in Figure 5 for the four conditions

332

investigated in this study, and it revealed that the ζ-potential is slightly negative following the

333

low-Mg2+ experiment. However, the particle surface charge decreased by 20 mV in the high-

334

Mg2+ experiment. This shift in ζ-potential might be a result of the decrease in pH that was

335

required to maintain equilibrium between the brucite and aqueous solution when MgCl2 was

336

added, which is expressed by eq 2. In both Mg2+ experiments, the ζ-potential indicated that the

337

negatively-charged brucite surface was electrostatically inhibited from adsorbing ribose, which

338

formed a surface complex as a negatively-charged species, Rib–. As the brucite surface became

339

more negative with increasing Mg2+ ion concentration, ribose adsorption further decreased,

340

which we have experimentally observed in the high-Mg2+ experiment.

341

In contrast to the Mg2+ experiments, we predicted that the ζ-potential significantly

342

increases toward positive values of 55 mV and 65 mV for the low- and high-Ca2+ experiments,

343

respectively. For both Ca2+ experiments, we observed a large amount of Ca adsorption, which

344

may have caused this predicted increase in the ζ-potential. Additionally, we observed a small

345

increase in ribose adsorption, which may have been promoted by the formation of the Ca-ribose

346

complex at the surface. This conclusion was supported by our predicted distribution of the ribose

347

species on brucite at low-Ca2+ and high-Ca2+ ion concentrations as shown by Figure 6 a,b. At

348

both Ca2+ ion concentrations, we predicted that the Ca-ribose outer-sphere species predominates

349

over the “standing” species on the brucite surface.

ACS Paragon Plus Environment

15

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 16 of 33

350

In our previous investigation of the brucite-aspartate system, we observed a significantly

351

greater increase in aspartate adsorption onto brucite under similar Ca2+ ion concentrations in

352

comparison to ribose35. The difference in the amount of adsorption between aspartate and ribose

353

with Ca highlights how a Ca-biomolecule surface complex is affected by different species. It is

354

possible that ribose forms a weaker attachment to Ca than the aspartate molecule, as is indicated

355

by the multiple bonding sites required to form the Ca-ribose complex and the low association

356

energy from the aqueous Ca-ribose complex51(Table 2). Should a positively-charged molecule

357

adsorb onto the brucite surface along with added CaCl2, it is possible that the molecule will

358

compete with the Ca2+ ion for adsorption onto the brucite surface. Such observations have been

359

previously reported in the case of lysine attachment with Ca2+ on rutile57 and montmorillonite58.

360

3.5. Prediction of Ribose Adsorption as a Function of pH

361

Once we were able to accurately model the dependence of ribose adsorption with ionic

362

strength and initial ribose concentration, we could also use the ETLM in a predictive manner to

363

extrapolate our results as a function of pH. Such a prediction is possible because in the course of

364

establishing a model for the brucite surface, we determined equilibrium constants for pH-

365

dependent surface protonation and deprotonation reactions (Table 2), as well as pH-dependent

366

surface reactions in eqs 4 and 5 and eqs 9 and 10. Because shifts in pH are accompanied by Mg2+

367

ion dissolution at the brucite surface, which maintains equilibrium with the aqueous phase (eq 2),

368

it is unlikely that these ionic strengths would naturally persist at this pH range. Nevertheless, this

369

prediction may provide a basic estimate of the magnitude of ribose adsorption in different

370

aqueous environments until further experiments can be done to better constrain the adsorption

371

predictions. As illustrated by Figure 7a,b, there is a significant amount of ribose surface

372

adsorption at low- and high- Ca2+ ion concentrations at pH 9 (Γads= 0.9 and 1.5 µmol•m-2 at

ACS Paragon Plus Environment

16

Page 17 of 33

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

373

[Rib]0 = 75 and 150 µM respectively). Additionally, we predict a similar extent of ribose

374

adsorption at both low- and high-Mg2+ ion concentrations under alkaline pH conditions.

375

Our predictions may indicate that ribose adsorption is favorable on brucite within alkaline

376

environments containing elevated Ca2+ ion concentrations. It may be possible that a

377

hyperalkaline aqueous system would act as a suitable environment for ribose adsorption,

378

although it is not likely that such an environment would be globally widespread. Nonetheless, the

379

condensation of ribose from simpler precursor molecules such as formaldehyde may favor

380

alkaline environments in addition to clay surfaces, including brucite15,33,59. Both brucite and

381

montmorillonite have been demonstrated to catalyze the formation of amino sugars and

382

pyrophosphate33,60,61, which are essential polymers for the formation of nucleic acids and the

383

evolution of RNA. Experimental observations of phosphorylation of nucleic acids on the mineral

384

schreibersite (Fe,Ni)3P have also indicated a prerequisite for alkaline conditions62. Our

385

predictions of increased ribose adsorption on brucite at elevated pH together with these previous

386

investigations may indicate that alkaline natural environments, such as serpentinizing

387

hydrothermal systems, may potentially favor the formation of ribose and the emergence of

388

ribonucleic acids at mineral surfaces.

389 390

4. Conclusions

391

We conducted low-Mg2+, high-Mg2+, low-Ca2+, and high-Ca2 batch adsorption

392

experiments to characterize the attachment of ribose onto brucite. Our results revealed that ribose

393

adsorption increased when we added CaCl2 to the system, whereas it decreased when we added

394

MgCl2. We characterized the brucite-ribose and brucite-Ca system with the ETLM and proposed

395

three types of surface reactions involving 1) a bidentate outer-sphere, or “standing” ribose

ACS Paragon Plus Environment

17

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 18 of 33

396

species, 2) a Ca-ribose complex attaching as a monodentate outer-sphere species, and 3) a

397

hydrated Ca complex attaching as a monodentate outer-sphere species. We tested our surface

398

complexation model against the data collected at four different ionic strengths, variable initial

399

ribose concentrations, and different pH conditions dependent on Mg2+ ion concentration. In each

400

case, the three surface reactions were consistent with the experimental data within analytical

401

uncertainty. Our success in representing the data with the ETLM has provided support that our

402

proposed surface reactions accurately depict ribose and Ca adsorption onto brucite.

403

We determined that ribose adsorption significantly decreases when MgCl2 is added, and

404

we predicted that the brucite particle surface charge becomes more negative with added MgCl2,

405

which creates an unfavorable electrostatic environment for the negatively-charged Rib- species to

406

adsorb as a “standing” complex. In contrast, we calculated a reversal in brucite particle surface

407

charge with the addition and subsequent adsorption of Ca, and the observed increase in ribose

408

adsorption may be due to the formation of a cooperative Ca-ribose complex. We predicted that

409

the Ca-ribose species predominates over the “standing” species at both Ca2+ ion concentrations.

410

However, we observed only a small increase in ribose surface adsorption when CaCl2 was added,

411

and this minor effect might be a consequence of the weak attachment of ribose to Ca.

412

We further used the ETLM to predict ribose adsorption over the pH range 5 to 12. The

413

model revealed that ribose surface adsorption might significantly increase at highly alkaline

414

conditions (Figure 7a,b), which might hold implications for ribose adsorption and concentration

415

at serpentinite-hosted hydrothermal environments. Arrhenius and coworkers50 have previously

416

observed that phosphate and polyphosphate intercalate into mixed-valence double layer

417

hydroxides (DLHs), in which trivalent cations Fe3+ or Al3+ have substituted into the brucite

418

structure. Given that the lateral brucite surface at which ribose sorbs is also located at the edge of

ACS Paragon Plus Environment

18

Page 19 of 33

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

419

the interlayer space of the brucite structure where the intercalation of phosphate is thought to

420

occur, it is possible that ribose might phosphorylate at a DLH surface containing phosphate.

421

Furthermore, a previous investigation50,51 has demonstrated that intercalated nucleic acids such

422

as AMP and even DNA can hybridize within the interlayer space of DLHs. The concentration of

423

ribose at the brucite (or DLH) surface at alkaline, CaCl2-rich conditions, as those found in

424

modern serpentinizing fluids, together with phosphate and nucleobases may influence the

425

assembly of ribonucleic acids, and possibly the emergence of RNA, through the process of

426

hybridization within the interlayer space. Our current model provides us with a fundamental

427

understanding of ribose adsorption behavior and the importance, as well as limitations, of

428

divalent ions in cooperatively enhancing its attachment onto the surface in the context of a

429

potentially more complex geochemical environment.

430 431

Supporting Information

432

Ribose polymerization blank experiment, surface complexation modeling considerations, SEM

433

image of brucite, equilibrium constant vs. pH of brucite, ribose and Ca adsorption data.

434 435

Acknowledgements

436

We thank Manuel Pelletier, Angélina Razafitianamaharavo, Cécile Feuillie, Namhey Lee,

437

George Cody, Timothy Strobel, Dionysis Foustoukos, Paul Goldey, John Armstrong, Adrian

438

Villegas-Jimenez, Stephen Hodge and Steven Coley for their invaluable expertise and advice

439

throughout this project. We also thank the three anonymous reviewers of this manuscript for

440

their useful comments and suggestions. This research was supported by the NASA Astrobiology

441

Institute, the Deep Carbon Observatory, Johns Hopkins University, the Carnegie Institute for

ACS Paragon Plus Environment

19

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 20 of 33

442

Science and funded by grants from the National Science Foundation EAR-1023865 (DAS) and

443

EAR-1023889 (RMH) and Department of Energy DE-FG02-96ER-14616.

444

References

445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484

(1)

(2)

(3) (4) (5)

(6)

(7)

(8)

(9) (10) (11) (12) (13)

(14) (15) (16) (17)

Berner, R. A.; Lasaga, A. C.; Garrels, R. M. The Carbonate-Silicate Geochemical Cycle and Its Effect on Atmospheric Carbon Dioxide Over the Past 100 Million Years. Am. J. Sci. 1983, 283, 641–683. Andrews, J. A.; Schlesinger, W. H. Soil CO2 Dynamics, Acidification, and Chemical Weathering in a Temperate Forest with Experimental CO2 Enrichment. Global Biogeochem. Cycles 2006, 15, 149–162. Schmitt-Kopplin, P.; Gabelica, Z.; Gougeon, R. D.; Fekete, A.; Kanawati, B.; Harir, M.; Gebefuegi, I.; Eckel, G.; Hertkorn, N. Proc. Natl. Acad. Sci. USA 2010, 107, 2763-2768. Stolz, J. F.; Chang, S.-B. R.; Kirschvink, J. L. Magnetotactic Bacteria and SingleDomain Magnetite in Hemipelagic Sediments. Nature 1986, 321, 849–851. Chan, C. S.; Fakra, S. C.; Emerson, D.; Fleming, E. J.; Edwards, K. J. Lithotrophic IronOxidizing Bacteria Produce Organic Stalks to Control Mineral Growth: Implications for Biosignature Formation. ISME J. 2010, 5 (4), 717–727. Williams, K. H.; Ntarlagiannis, D.; Slater, L. D.; Dohnalkova, A.; Hubbard, S. S.; Banfield, J. F. Geophysical Imaging of Stimulated Microbial Biomineralization. Environ. Sci. Technol. 2005, 39 (19), 7592–7600. Brasier, M. D.; Antcliffe, J.; Saunders, M.; Wacey, D. Changing the Picture of Earth's Earliest Fossils (3.5–1.9 Ga) with New Approaches and New Discoveries. Proc. Natl. Acad. Sci. USA 2015, 112 (16), 4859–4864. Hazen, R. M. Presidential Address to the Mineralogical Society of America, Salt Lake City, October 18, 2005: Mineral Surfaces and the Prebiotic Selection and Organization of Biomolecules. Am. Mineral. 2006, 91 (11-12), 1715–1729. Hazen, R. M.; Sverjensky, D. A. Mineral Surfaces, Geochemical Complexities, and the Origins of Life. Cold Spring Harb. Perspect. Biol. 2010, 2 (5), a002162. Hazen, R. M. Genesis: Rocks, Minerals, and the Geochemical Origin of Life. Elements 2005, 1 (3), 135–137. Ferris, J. P. Catalysis and Prebiotic RNA Synthesis. Orig. Life Evol. Biosph. 1993, 23 (56), 307–315. Huang, W.; Ferris, J. P. One-Step, Regioselective Synthesis of Up to 50-Mers of RNA Oligomers by Montmorillonite Catalysis. J. Am. Chem. Soc. 2006, 128 (27), 8914–8919. Malin, J. N.; Holland, J. G.; Geiger, F. M. Free Energy Relationships in the Electric Double Layer and Alkali Earth Speciation at the Fused Silica/Water Interface. J. Phys. Chem. C 2009, 113 (41), 17795–17802. Angyal, S. J. The Composition and Conformation of Sugars in Solution. Angew.. Chem. Int. Ed. Engl. 1969, 8 (3), 157–166. Gabel, N. W.; Ponnamperuma, C. Model for Origin of Monosaccharides. Nature 1967, 216 (5114), 453–455. Reid, C.; Orgel, L. E.; Ponnamperuma, C. Nucleoside Synthesis Under Potentially Prebiotic Conditions. Nature 1967, 216 (5118), 936–936. Kim, H.-J.; Ricardo, A.; Illangkoon, H. I.; Kim, M. J.; Carrigan, M. A.; Frye, F.; Benner,

ACS Paragon Plus Environment

20

Page 21 of 33

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

485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530

ACS Earth and Space Chemistry

(18)

(19)

(20)

(21)

(22)

(23)

(24) (25)

(26)

(27) (28)

(29) (30)

(31)

(32)

(33)

S. A. Synthesis of Carbohydrates in Mineral-Guided Prebiotic Cycles. J. Am. Chem. Soc. 2011, 133 (24), 9457–9468. LaRowe, D. E.; Regnier, P. Thermodynamic Potential for the Abiotic Synthesis of Adenine, Cytosine, Guanine, Thymine, Uracil, Ribose, and Deoxyribose in Hydrothermal Systems. Orig. Life Evol. Biosph. 2008, 38 (5), 383–397. Charlou, J. L.; Donval, J. P.; Fouquet, Y.; Jean-Baptiste, P.; Holm, N. G. Geochemistry of High H2 and CH4 Vent Fluids Issuing From Ultramafic Rocks at the Rainbow Hydrothermal Field (36 14'N, MAR). Chem. Geol. 2002, 191 (4), 345–359. Delacour, A. L.; Früh-Green, G. L.; Bernasconi, S. M.; Schaeffer, P.; Kelley, D. S. Carbon Geochemistry of Serpentinites in the Lost City Hydrothermal System (30°N, MAR). Geochim. Cosmochim. Acta 2008, 72 (15), 3681–3702. Feuillie, C.; Daniel, I.; Michot, L. J.; Pedreira-Segade, U. Adsorption of Nucleotides Onto Fe–Mg–Al Rich Swelling Clays. Geochim. Cosmochim. Acta 2013, 120 (C), 97– 108. Banin, A.; Lawless, J. G.; Mazzurco, J.; Church, F. M.; Margulies, L.; Orenberg, J. B. pH Profile of the Adsorption of Nucleotides Onto Montmorillonite. Orig. Life Evol. Biosph. 1985, 15 (2), 89–101. Hashizume, H.; van der Gaast, S.; Theng, B. K. G. Adsorption of Adenine, Cytosine, Uracil, Ribose, and Phosphate by Mg-Exchanged Montmorillonite. Clay Miner. 2010, 45 (4), 469–475. Rishpon, J.; O'Hara, P. J.; Lahav, N.; Lawless, J. G. Interaction Between ATP, Metal Ions, Glycine, and Several Minerals. J. Mol. Evol. 1982, 18 (3), 179–184. Feuillie, C.; Sverjensky, D. A.; Hazen, R. M. Attachment of Ribonucleotides on ΑAlumina as a Function of pH, Ionic Strength, and Surface Loading. Langmuir 2015, 31 (1), 240–248. Holland, J. G.; Malin, J. N.; Jordan, D. S.; Geiger, F. M. Specific and Nonspecific Metal Ion−Nucleotide Interactions at Aqueous/Solid Interfaces Functionalized with Adenine, Thymine, Guanine, and Cytosine Oligomers. J. Am. Chem. Soc. 2011, 133 (8), 2567– 2570. Arora, A. K.; Kamaluddin. Interaction of Ribose Nucleotides with Zinc Oxide and Relevance in Chemical Evolution. Colloids Surf., A 2007, 298 (3), 186–191. Bhushan, B.; Shanker, U.; Kamaluddin. Adsorption of Ribose Nucleotides on Manganese Oxides with Varied Mn/O Ratio: Implications for Chemical Evolution. Orig. Life Evol. Biosph. 2011, 41 (5), 469–482. Hashizume, H.; Theng, B. K. G. Adenine, Adenosine, Ribose and 5′-Amp Adsorption to Allophane. Clay Clay Miner. 2007, 55 (6), 599–605. Cleaves, H. J., II; Jonsson, C. M.; Jonsson, C. L.; Sverjensky, D. A.; Hazen, R. M. Adsorption of Nucleic Acid Components on Rutile (TiO2) Surfaces. Astrobiology 2010, 10 (3), 311–323. Kelley, D. S.; Karson, J. A.; Früh-Green, G. L.; Yoerger, D. R.; Shank, T.; Butterfield, D. A.; Hayes, J.; Schrenk, M. O.; Olson, E. J.; Proskurowski, G. A Serpentinite-Hosted Ecosystem: the Lost City Hydrothermal Field. Science 2005, 307 (5714), 1428. Proskurowski, G.; Lilley, M. D.; Kelley, D. S.; Olson, E. J. Low Temperature Volatile Production at the Lost City Hydrothermal Field, Evidence From a Hydrogen Stable Isotope Geothermometer. Chem. Geol. 2006, 229 (4), 331–343. Holm, N. G.; Dumont, M.; Ivarsson, M.; Konn, C. Alkaline Fluid Circulation in

ACS Paragon Plus Environment

21

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

531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576

(34) (35)

(36)

(37)

(38)

(39)

(40) (41)

(42)

(43)

(44)

(45)

(46)

(47) (48)

Page 22 of 33

Ultramafic Rocks and Formation of Nucleotide Constituents: a Hypothesis. Geochem Trans 2006, 7 (1), 1–13. Holm, N. G. The Significance of Mg in Prebiotic Geochemistry. Geobiology 2012, 10 (4), 269–279. Estrada, C. F.; Sverjensky, D. A.; Pelletier, M.; Razafitianamaharavo, A.; Hazen, R. M. Interaction Between L-Aspartate and the Brucite [Mg(OH)2]–Water Interface. Geochim. Cosmochim. Acta 2015, 155, 172–186. Henrist, C.; Mathieu, J.; Vogels, C.; Rulmont, A.; Cloots, R. Morphological Study of Magnesium Hydroxide Nanoparticles Precipitated in Dilute Aqueous Solution. J. Cryst. Growth 2003, 249, 321–330. Lu, J.; Qiu, L.; Qu, B. Controlled Growth of Three Morphological Structures of Magnesium Hydroxide Nanoparticles by Wet Precipitation Method. J. Cryst. Growth 2004, 267 (3-4), 676–684. Villiéras, F.; Cases, J.-M.; François, M.; Michot, L. J.; Thomas, F. Texture and Surface Energetic Heterogeneity of Solids From Modeling of Low Pressure Gas Adsorption Isotherms. Langmuir 1992, 8 (7), 1789–1795. Villiéras, F.; Michot, L. J.; Bardot, F.; Cases, J.-M.; François, M.; Rudziński, W. An Improved Derivative Isotherm Summation Method to Study Surface Heterogeneity of Clay Minerals. Langmuir 1997, 13 (5), 1104–1117. Michot, L. J.; Villiéras, F. Assessment of Surface Energetic Heterogeneity of Synthetic Na-Saponites. the Role of Layer Charge. Clay Miner. 2002, 37 (1), 39–57. Eypert-Blaison, C.; Villiéras, F.; Michot, L. J.; Pelletier, M.; Humbert, B.; Ghanbaja, J.; Yvon, J. Surface Heterogeneity of Kanemite, Magadiite and Kenyaite: a HighResolution Gas Adsorption Study. Clay Miner. 2002, 37 (3), 531–542. Sayed-Hassan, M.; Villiéras, F.; Gaboriaud, F.; Razafitianamaharavo, A. AFM and LowPressure Argon Adsorption Analysis of Geometrical Properties of Phyllosilicates. J. Colloid Interface Sci. 2006, 296 (2), 614–623. Perronnet, M.; Villiéras, F.; Jullien, M.; Razafitianamaharavo, A.; Raynal, J.; Bonnin, D. Towards a Link Between the Energetic Heterogeneities of the Edge Faces of Smectites and Their Stability in the Context of Metallic Corrosion. Geochim. Cosmochim. Acta 2007, 71 (6), 1463–1479. Pokrovsky, O. S.; Schott, J.; Castillo, A. Kinetics of Brucite Dissolution at 25°C in the Presence of Organic and Inorganic Ligands and Divalent Metals. Geochim. Cosmochim. Acta 2005, 69 (4), 905–918. Jandik, P.; Pohl, C. A.; Barreto, V.; Avdalovic, N. Anion Exchange Chromatography and Integrated Amperometric Detection of Amino Acids. In Amino Acid Analysis Protocols; Humana Press: New Jersey, 2000; Vol. 159, pp 63–85. Clarke, A. P.; Jandik, P.; Rocklin, R. D.; Liu, Y.; Avdalovic, N. An Integrated Amperometry Waveform for the Direct, Sensitive Detection of Amino Acids and Amino Sugars Following Anion-Exchange Chromatography. Anal. Chem. 1999, 71 (14), 2774– 2781. Jensen, D.; Weiss, J.; Rey, M. A.; Pohl, C. A. Novel Weak Acid Cation-Exchange Column. J. Chromatogr., A 1993, 640 (1-2), 65–71. Sverjensky, D. A.; Fukushi, K. Anion Adsorption on Oxide Surfaces: Inclusion of the Water Dipole in Modeling the Electrostatics of Ligand Exchange. Environ. Sci. Technol. 2006, 40 (1), 263–271.

ACS Paragon Plus Environment

22

Page 23 of 33

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

577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621

ACS Earth and Space Chemistry

(49) (50)

(51)

(52)

(53)

(54)

(55) (56)

(57)

(58)

(59) (60)

(61)

(62)

(50) (51)

Sahai, N.; Sverjensky, D. A. GEOSURF: a Computer Program for Modeling Adsorption on Mineral Surfaces From Aqueous Solution. Comput. Geosci. 1998, 24 (9), 853–873. Goldberg, R. N.; Tewari, Y. B. Thermodynamic and Transport-Properties of Carbohydrates and Their Monophosphates - the Pentoses and Hexoses. J. Phys. Chem. Ref. Data 1989, 18 (2), 809–880. Alvarez, A. M.; Morel-Desrosiers, N.; Morel, J.-P. Interactions Between Cations and Sugars. III. Free Energies, Enthalpies, and Entropies of Association of Ca2+, Sr2+, Ba2+, La3+, Gd3+ with D-Ribose in Water at 25° C. Can. J. Chem. 1987, 65, 2656–2660. Pokrovsky, O. S.; Schott, J. Experimental Study of Brucite Dissolution and Precipitation in Aqueous Solutions: Surface Speciation and Chemical Affinity Control. Geochim. Cosmochim. Acta 2004, 68 (1), 31–45. Prélot, B.; Villiéras, F.; Pelletier, M.; Gérard, G.; Gaboriaud, F.; Ehrhardt, J.-J.; Perrone, J.; Fedoroff, M.; Jeanjean, J.; Lefèvre, G.; et al. Morphology and Surface Heterogeneities in Synthetic Goethites. J. Colloid Interface Sci. 2003, 261 (2), 244–254. Liu, X.; Cheng, J.; Sprik, M.; Lu, X.; Wang, R. Understanding Surface Acidity of Gibbsite with First Principles Molecular Dynamics Simulations. Geochim. Cosmochim. Acta 2013, 120 (C), 487–495. Sverjensky, D. A. Standard States for the Activities of Mineral Surface Sites and Species. Geochim. Cosmochim. Acta 2003, 67 (1), 17–28. Lu, Y.; Deng, G.; Miao, F.; Li, Z. Sugar Complexation with Calcium Ion. Crystal Structure and FT-IR Study of a Hydrated Calcium Chloride Complex of D-Ribose. J. Inorg. Biochem. 2003. Lee, N.; Sverjensky, D. A.; Hazen, R. M. Cooperative and Competitive Adsorption of Amino Acids with Ca2+ on Rutile (α-TiO2). Environ. Sci. Technol. 2014, 48 (16), 9358– 9365. Yang, Y.; Wang, S.; Liu, J.; Xu, Y.; Zhou, X. Adsorption of Lysine on NaMontmorillonite and Competition with Ca2+: a Combined XRD and ATR-FTIR Study. Langmuir 2016, 32 (19), 4746–4754. Zubay, G. Studies on the Lead-Catalyzed Synthesis of Aldopentoses. Orig. Life Evol. Biosph. 1998, 28, 13–26. Hermes-Lima, M.; Vieyra, A. Pyrophosphate Formation From Phospho(Enol)Pyruvate Adsorbed Onto Precipitated Orthophosphate: a Model for Prebiotic Catalysis of Transphosphorylations. Orig. Life Evol. Biosph. 1989, 19, 143–152. Hermes-Lima, M.; Vieyra, A. Pyrophosphate Synthesis From Phospho(Enol)Pyruvate Catalyzed by Precipitated Magnesium Phosphate with “Enzyme-Like” Activity. J. Mol. Evol. 35, 277–285. Gull, M.; Mojica, M. A.; Fernández, F. M.; Gaul, D. A.; Orlando, T. M.; Liotta, C. L.; Pasek, M. A. Nucleoside Phosphorylation by the Mineral Schreibersite. Scientific Reports 2017, 1–6. Arrhenius, G.; Sales, B.; Mojzsis, S.; Lee, T. Entropy and Charge in Molecular Evolution—the Case of Phosphate. J. Theor. Biol. 1997, 187 (4) 503-522. Choy, J-H.; Kwak, S. Y.; Park, J. S.; Jeong Y. J. Intercalative Nanohybrids of Nucleoside Monophosphates and DNA in Layered Metal Hydroxide. J. Am. Chem. Soc. 1999, 121 (6), 1399-1400.

ACS Paragon Plus Environment

23

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

Page 24 of 33

Figure 1. Adsorption of ribose on brucite as function of the concentration remaining in water for the (a) low-Mg2+ (red) and high-Mg2+ (blue) experiments and the (b) low-Ca2+ (purple) and high-Ca2+ (green) experiments. Predictions of ribose adsorption from the surface reactions in eqs 4 and 5 are indicated by dashed and dotted lines, respectively. Symbols represent experimental data following a triplicate run and vertical error bars represent one standard deviation.

ACS Paragon Plus Environment

24

Page 25 of 33

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

ACS Earth and Space Chemistry

Figure 2. Two potential visual representations of the ribose surface species predicted from eqs 4 and 5 as a bidentate outer-sphere, or “standing” species using the ETLM and the parameters in Table 2. Ribose may attach onto (a) one neutral and one positively-charged surface site or (b) two positively-charged surface sites. The idealized, lateral (100) brucite growth surface is shown with red spheres as O, yellow spheres as Mg, light pink spheres as H, and black spheres as C atoms.

ACS Paragon Plus Environment

25

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

Page 26 of 33

Figure 3. Adsorption of Ca onto brucite as a function of initial ribose concentration added for the low-Ca2+ experiment (purple) and the high-Ca2+ experiment (green). Predictions of Ca adsorption from the surface reactions in eqs 9 and 10 are indicated by dashed curves. Symbols represent experimental data following a triplicate run and vertical error bars represent one standard deviation.

ACS Paragon Plus Environment

26

Page 27 of 33

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

ACS Earth and Space Chemistry

Figure 4. A visual representation of the (a) cooperative Ca-ribose and (b) a hydrated Ca surface outer-sphere species predicted by eqs 9 and 10 using the ETLM and the parameters in Table 2. The idealized, lateral (100) brucite surface is shown with red spheres as O, yellow spheres as Mg, light pink spheres as H, black spheres as C atoms, and light blue spheres as Ca.

ACS Paragon Plus Environment

27

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

Page 28 of 33

Figure 5. ζ-potential (mv), or particle surface charge, predicted with the ETLM and the parameters in Table 2 as a function of initial ribose concentration and the four experimental conditions examined in this study. Predictions from eqs 4 and 5 are represented with dashed and dotted curves, respectively, and overlay one another for the much of studied initial ribose conditions.

ACS Paragon Plus Environment

28

Page 29 of 33

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

ACS Earth and Space Chemistry

Figure 6. Predicted surface speciation of ribose on brucite for the (a) low-Ca2+ and (b) high-Ca2+ batch adsorption experiments. The “standing” (blue) and Ca outer-sphere (orange) species refer to those proposed in eqs 4 and 5, and eq 9, respectively. Predictions from eq 4 (dashed curve) and eq 5 (dotted curve) overlay one another over the entire initial ribose concentration range. Symbols represent experimental data following a triplicate run and vertical error bars represent one standard deviation.

ACS Paragon Plus Environment

29

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

Page 30 of 33

Figure 7. Predictions of ribose adsorption on brucite as a function of pH at the low-Mg2+ (red), high-Mg2+ (blue), low-Ca2+ (purple), and high-Ca2+ (green) conditions at (a) 75 µM and (b) 150 µM ribose. Predictions are based on existing adsorption data (symbols), ETLM parameters in Table 2, and eqs 4 and 5, which are indicated by the dashed and dotted curves, respectively.

ACS Paragon Plus Environment

30

Page 31 of 33

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

ACS Earth and Space Chemistry

Table 1. Description of the low-Mg2+, high-Mg2+, low-Ca2+, and high-Ca2+ experiments. Dissolved Mg2+ concentrations from brucite are present in each experiment in addition to added MgCl2 and CaCl2. All concentrations are expressed in mmol/L.

Experiment

Dissolved Mg2+

Added MgCl2

Added CaCl2

Ionic Strength

pH

Low-Mg2+

0.6

--

--

2.1

10.4

High-Mg2+

0.6

5.2

--

17.4

9.8

Low-Ca2+

0.9

--

1.0

5.8

10.2

High-Ca2+

0.6

--

4.2

15.2

10.3

ACS Paragon Plus Environment

31

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

Page 32 of 33

Table 2. Aqueous ribose properties, characteristics of brucite [Mg(OH2)], and extended triple-layer model parameters for proton electrolyte and ribose adsorption onto brucite. Reaction Type Aqueous ribose equilibria Surface equilibria

Reaction Rib– + H+ = HRib0

log K 12.00

HRib0 + Ca2+ = HRibCa2+

0.3051

Hypothetical 1.0 m standard state

logK=$

>SOH + H+ = >SOH2+

9.69

logK $6

>SO- + H+ = >SOH

11.31

log ∗ K $>

>SOH + Na+ = >SO-_Na+ + H+

-8.91

log ∗ K $?#

>SOH + Cl- + H+ = >SOH2+_Cl-

11.89

2>SOH + HRib0 = >SOH>SOH2+_Rib–

4.97

log ∗ K $_ 

log ∗ K $6_ 

log ∗ K $_

!"#

!"#

2+

!"

log ∗ K $__() 

2>SOH + H+ + HRib0 = 2>SOH2+_Rib– 2+

logK @_

+

-3.43 -5.11

2>SOH2 + HRib0 = >SOH>SOH2+_Rib– + 2H+

7.51

>SOH + Ca + 2H2O = >SOH2 _Ca(OH)2 + H Site-occupancy standard states

logK @_ 

15.45 + +

Surface equilibria

logK @6_

+

>SOH + Ca + HRib = >SOH_CaRib + H

a



0

+

!"#

+

!"#

0

logK @__()



+

2>SOH2 + HRib = 2>SOH2 _Rib + H 2+

!"

+

0

+

17.99 +

>SOH + Ca + HRib = >SOH_CaRib + H +

2+

+

>SOH2 + Ca + 2H2O = >SOH2 _Ca(OH2) + H

-3.66 +

-5.34



a

Equilibrium constants relative to site occupancy standard states can be written relative to charged surface sites calculated using the equations: logK @ _ !"# = log ∗ K $ _ !"# + log(NsAsCs)/100 ; logK @6 _ !"# = log ∗ K $6 _ !"# + log(NsAsCs)/100; logK @_ !" = 







log ∗ K $_ !" + log(NsAs)/100 ; logK @ __() = log ∗ K $ __() + log(NsAs)/100, where Ns is site density, As is BET edge surface area m2•g-1, and     Cs is the mineral loading g•L-1.

ACS Paragon Plus Environment

32

Page 33 of 33

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

ACS Earth and Space Chemistry

Graphical Abstract

ACS Paragon Plus Environment

33