Kinetics of α-MnOOH Nanoparticle Formation ... - ACS Publications

Sep 6, 2017 - Department of Energy, Environmental and Chemical Engineering, Washington University, St. Louis, Missouri 63130, United States. •S Supp...
2 downloads 12 Views 769KB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Kinetics of #-MnOOH Nanoparticle Formation through Enzymatically-catalyzed Biomineralization inside Apoferritin Yue Hui, Haesung Jung, Doyoon Kim, and Young-Shin Jun Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.7b00568 • Publication Date (Web): 06 Sep 2017 Downloaded from http://pubs.acs.org on September 9, 2017

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

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

Page 1 of 34

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

Crystal Growth & Design

Kinetics of -MnOOH Nanoparticle Formation through Enzymatically-catalyzed Biomineralization inside Apoferritin Yue Hui,† Haesung Jung,† Doyoon Kim, and Young-Shin Jun* Department of Energy, Environmental and Chemical Engineering, Washington University, St. Louis, Missouri 63130, United States †

These authors contributed equally *To whom correspondence should be addressed

Abstract While biomineralization in apoferritin has effectively synthesized highly monodispersed nanoparticles of various metal oxides and hydroxides, the detailed kinetics and mechanisms of Mn(III) (hydr)oxide formation inside apoferritin cavities have not been reported. To address this knowledge gap, we first identified the phase of solid Mn(III) formed inside apoferritin cavities as -MnOOH. To analyze the oxidation and nucleation mechanism of -MnOOH inside apoferritin by quantifying oxidized Mn, we used a colorimetric method with leucoberbelin blue (LBB) solution. In this method, LBB dissembled apoferritin by inducing an acidic pH environment, and reduced MnOOH nanoparticles. The LBB-enabled kinetic analyses of -MnOOH nanoparticle formation suggested that the orders of reaction with respect to Mn2+ and OH- are 2 and 4, respectively, and -MnOOH formation follows two-step pathways: First, soluble Mn2+ undergoes apoferritin catalyzed oxidation at the ferroxidase dinuclear center, forming a Mn(III)-protein complex, P[Mn2O2(OH)2]. Second, the oxidized Mn(III) dissociates from the protein binding sites and are subsequently nucleated to form -MnOOH nanoparticles in the apoferritin cavities. This study reveals key kinetics and mechanistic information of the Mn-apoferritin systems and the results facilitate applications of apoferritin as a means of nanomaterial synthesis.  

Young-Shin Jun, Ph.D., Professor Department of Energy, Environmental and Chemical Engineering, Washington University, St. Louis, MO 63130 Phone: (314)-935-4539 Fax: (314)-935-7211 E-mail: [email protected]

ACS Paragon Plus Environment

Crystal Growth & Design

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

Kinetics of -MnOOH Nanoparticle Formation through Enzymatically-catalyzed Biomineralization inside Apoferritin

Yue Hui,† Haesung Jung,† Doyoon Kim, and Young-Shin Jun*

Department of Energy, Environmental and Chemical Engineering, Washington University, St. Louis, MO 63130

E-mail: [email protected] http://encl.engineering.wustl.edu/ Submitted: August 2017

Crystal Growth & Design



These authors contributed equally

*To whom correspondence should be addressed

1

ACS Paragon Plus Environment

Page 2 of 34

Page 3 of 34

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

Crystal Growth & Design

1

ABSTRACT

2

While biomineralization in apoferritin has effectively synthesized highly monodispersed

3

nanoparticles of various metal oxides and hydroxides, the detailed kinetics and

4

mechanisms of Mn(III) (hydr)oxide formation inside apoferritin cavities have not been

5

reported. To address this knowledge gap, we first identified the phase of solid Mn(III)

6

formed inside apoferritin cavities as -MnOOH. To analyze the oxidation and nucleation

7

mechanism of -MnOOH inside apoferritin by quantifying oxidized Mn, we used a

8

colorimetric method with leucoberbelin blue (LBB) solution. In this method, LBB

9

dissembled apoferritin by inducing an acidic pH environment, and reduced -MnOOH

10

nanoparticles. The LBB-enabled kinetic analyses of -MnOOH nanoparticle formation

11

suggested that the orders of reaction with respect to Mn2+ and OH- are 2 and 4, respectively,

12

and -MnOOH formation follows two-step pathways: First, soluble Mn2+ undergoes

13

apoferritin catalyzed oxidation at the ferroxidase dinuclear center, Mn(III)-protein

14

complex, P-[Mn2O2(OH)2]. Second, the oxidized Mn(III) dissociates from the protein

15

binding sites and are subsequently nucleated to form -MnOOH nanoparticles in the

16

apoferritin cavities. This study reveals key kinetics and mechanistic information of the Mn-

17

apoferritin systems and the results facilitate applications of apoferritin as a means of

18

nanomaterial synthesis.

ACS Paragon Plus Environment

Crystal Growth & Design

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

19

Introduction

20

Self-assembled protein cages and viral capsids have gained wide interest from

21

materials scientists. Because of their shell-like structures with hollow interior space for

22

bio-mineralization, they can serve as nanoreactors in the synthesis of a variety of highly

23

monodispersed, inorganic nanomaterials.1-8 The variety of protein cage architectures

24

allows precise control of the sizes and shapes of particles without using toxic organic

25

surfactants, and results in nanoparticles with superior biocompatibility and new

26

functionalities.9-12 Given these apparent advantages, extensive efforts have been made to

27

establish a library of nanoreactors for synthesis of nanomaterials with specific properties.1,

28

3-6, 13-17

29

Cornelissen et al. generated monodisperse Prussian blue particles with a diameter of 18 ±

30

1.7 nm through photocatalysis.18 These synthesized protein-virus bio-hybrids were found

31

to self-organize in monolayer fashion on mica and other hydrophilic graphite surfaces.18

32

In addition, Watanabe et al. reported the synthesis of Pd nanoclusters in ferritin cavities

33

through reduction of Pd2+ by NaBH4.19 Compared to Pd particles alone, the Pd-

34

encapsulated ferritin demonstrated a superior ability to catalyze size-selective

35

hydrogenation of olefins.

For example, utilizing cowpea chlorotic mottle virus (CCMV) as a cage template,

36

Ferritin, one of the most studied protein cages for nanoparticle synthesis, comprises

37

a class of iron storage proteins ubiquitously present among living species.20, 21 Ferritin

38

serves as an iron reservoir by sequestering iron when the cellular iron supply is high.20-22

39

With its architecture of a hollow spherical shell with an outer diameter of 12 nm and a

40

central cavity 8 nm in diameter, a ferritin molecule can store up to 4500 iron atoms in the

41

form of Fe(III) (oxy)hydroxide through sequential oxidation and mineralization

2

ACS Paragon Plus Environment

Page 4 of 34

Page 5 of 34

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

Crystal Growth & Design

42

processes.20, 21 Ferritin is composed of 24 subunits of two types, heavy (H) and light (L).23-

43

25

44

in catalyzing the oxidation of Fe2+. Although lacking the catalytic site for oxidation, the L

45

subunit is more efficient in the subsequent mineralization of Fe(III) because of the strong

46

negative charges carried by the glutamate clusters that serve as the nucleation sites for

47

mineral formation.23-25 Kinetic studies of Fe(III) (hydr)oxide formation inside apoferritin

48

have revealed two iron concentration-dependent mechanisms for the overall iron oxidation

49

and nucleation processes.20, 26-29 Under low iron influx (less than 48 Fe(II) atoms/protein),

50

oxidation of Fe2+ to Fe3+ occurs predominantly at the ferroxidase center, where the

51

enzymatic sites bind two Fe(II), and hence catalyze their oxidation in the reaction with

52

dissolved oxygen, producing H2O2. As iron flux increases (more than 48 Fe(II)

53

atoms/protein) and ferroxidase sites become saturated with Fe(II), Fe(II) oxidation and

54

sequential Fe(III) hydrolysis occur directly on the pre-nucleated Fe(III) (oxy)hydroxide

55

surface.20, 26-29

The H subunit, with a conserved dinuclear ferroxidase center, exhibits enzymatic activity

56

Since the mechanism of iron deposition in native ferritin has been revealed, many

57

studies have attempted to use the empty ferritin shell (apoferritin) as a macromolecular

58

template for synthesis of a variety of inorganic nanoparticles, including oxides or

59

hydroxides of cobalt,30 chromium,16 nickel,16 and indium,31 cadmium selenide,32 and

60

cobalt/platinum alloys.33 The synthesized particles, together with the protein cage, have

61

been employed in a variety of biomedical applications, such as drug delivery.34-39 The

62

synthesis of Mn-apoferritin nanocomposite, in particular, has attracted great interest from

63

the materials science field because of its use as a highly-functioning contrast agent in

64

MRI.40-42 Mann’s group was the first to observe the successful synthesis of MnOOH

3

ACS Paragon Plus Environment

Crystal Growth & Design

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

65

nanoparticles in a ferritin cavity.43, 44 To elucidate the effect of apoferritin composition on

66

MnOOH synthetic process, in their studies, they investigated recombinant apoferritin

67

molecules with varied proportions of H- and L-chain subunits. In making their successful

68

observations of MnOOH formation in apoferritin, Mann’s group correlated spectroscopic

69

measurement at 450 nm with Mn(III) (oxy) hydroxide core formation inside apoferritin. In

70

their study, the concentrations of solid Mn(III) formed inside apoferritin cavities could not

71

be specified, and thus the detailed kinetics and mechanism of Mn(II) oxidation and Mn(III)

72

(oxy) hydroxide nucleation in apoferritin were not reported. To better control of material’s

73

properties, however, it is important to have sufficient knowledge about the mechanisms

74

governing the protein-manganese interaction. Otherwise, the biomineralization process has

75

remained largely uncontrolled, hence hindering its application in areas where time-wise

76

particle size and morphology control are required.

77

In addition, while previous studies on the Fe-apoferritin system have revealed

78

detailed mechanisms of oxidation and mineralization of FeOOH through kinetic analyses,

79

the differences between the reaction pathways of the Fe and Mn systems remain largely

80

unexplored.27,28,39 Thus, it is unclear whether the mechanism in Fe-apoferritin systems can

81

sufficiently explain that in Mn-apoferritin systems. Therefore, the purpose of our study is

82

to investigate Mn-apoferritin systems, determining the stoichiometry and elucidating the

83

step-wise reaction pathways for Mn oxidation and nucleation processes inside apoferritin

84

cavities. In this study, we confirmed the specific phase of Mn(III) (hydr)oxide and provided

85

detailed kinetic analyses of MnOOH formation inside apoferritin cavities by quantifying

86

solid Mn(III) concentrations with the leucoberbelin blue (LBB) method. We also

87

investigated the kinetics of -MnOOH formation. From there, we elucidated the

4

ACS Paragon Plus Environment

Page 6 of 34

Page 7 of 34

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

Crystal Growth & Design

88

similarities and differences in the oxidation mechanisms involved in Fe- and Mn-

89

apoferritin systems. Our findings provide a new understanding of the functionality of

90

apoferritin for MnOOH synthesis and allow more systematic prediction and control of

91

particle sizes and morphology when using macromolecular templates for synthesizing

92

metal (hydr)oxide nanomaterials.

93

Experimental Section

94

Sample preparation

95

Sample solutions were prepared using apoferritin extracted from equine spleen (0.2

96

µm filtered, Sigma-Aldrich, consisting of around 10% H and 90% L subunits45-47), reagent-

97

grade Mn(NO3)2·4H2O (99.98%, Alfa Aesar), AMPSO (99%, Sigma-Aldrich), and aerated

98

ultrapure deionized water (resistivity >18.2 MΩ-cm, with 8.4 ± 0.1 mg/L of dissolved O2).

99

All samples were prepared in 10 mL 0.05 M AMPSO buffer solution with 0.1 µM

100

apoferritin. This preparation allowed us to maintain an initial pH condition during an

101

experiment lasting several hours without any changes of the initial concentrations of Mn2+

102

(aq) and apoferritin and without precipitation of Mn(OH)2 (s)48, 49, all of which can occur

103

if pH is maintained with dynamic titration. The AMPSO buffer has been widely employed

104

in previous studies on MnOOH formation inside apoferritin, and thus we consider the use

105

of AMPSO to be a proper experimental approach that can be more comparable to those in

106

previous literature.43, 44

107

To examine the effect of OH- concentration on the rate of MnOOH formation, pH

108

conditions of 8.90 ± 0.05, 9.00 ± 0.05, and 9.10 ± 0.05 were investigated. These pH ranges

109

were chosen because it enables measuring the nucleation kinetics of MnOOH in apoferritin

110

within hours, and can differentiate the nucleation of MnOOH in apoferritin and in solution.

5

ACS Paragon Plus Environment

Crystal Growth & Design

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

111

All samples used for Mn(III) quantification were replicated more than three times to

112

account for the high sensitivity of the -MnOOH formation rate in apoferritin to changes

113

in pH. To reach the target pH values, proper amounts of 1 M HNO3 solution were added.

114

For each pH condition, samples with different Mn2+ concentrations were prepared by

115

adjusting the initial Mn2+ atom: apoferritin ratios to be 2000:1, 2500:1, 3000:1, 3500:1, and

116

4000:1.

117

Samples in test tubes were parafilm-sealed with a small punched hole to allow free

118

access to air but to prevent unwanted contamination or significant water evaporation.

119

Because oxygen was constantly replenished due to contact with the atmosphere and the

120

sample solutions remained equilibrated with oxygen, we assumed that oxygen

121

consumption was not a limiting factor for Mn oxidation kinetics. The measured oxygen

122

concentration at pH 9.1 with an Mn to apoferritin ratio of 4000:1, which shows the fastest

123

oxidation rate among the experimental conditions, also supports considering the oxygen

124

concentration as a constant value (Figure S1). To analyze the difference in nucleation

125

pathways in systems with and without apoferritin, control experiments were conducted

126

without the addition of apoferritin.

127

Phase characterization

128

The oxidation states of the Mn solid phase in samples prepared with and without

129

apoferritin were characterized using X-ray Photoelectron Spectroscopy (XPS,

130

PHI 5000 VersaProbe II, Ulvac-PHI with monochromatic Al Kα radiation (1486.6 eV)).

131

The Mn 3p spin orbit was used because it provides more accurate information than the Mn

132

2p that has less sensitivity to the bonding environment of the mineral.50 Likewise, because

133

the intensity of Mn 3s spin orbit was too weak to get information about multiplet splitting,

6

ACS Paragon Plus Environment

Page 8 of 34

Page 9 of 34

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

Crystal Growth & Design

134

we used only Mn 3p spin orbit. From this analysis, the peak positions of experimental

135

samples were compared to those of reference samples of Mn(II) and Mn(III). The peak

136

positions of Mn(II) and Mn(III) used for reference were 48.1 eV and 48.7 eV, based on the

137

average of values reported in previous works and those of purchased samples from the

138

natural environment (S2 in the Supporting Information). C 1s at 284.8 eV was used as the

139

reference energy level. To determine the oxidation states of Mn, peak positions given by

140

the asymmetrical Gaussian−Lorentzian curve-fitting for the experimental samples were

141

compared to the reference peak values.

142

High resolution transmission electron microscopy (HRTEM, JEOL 2100F) images

143

of reaction systems with and without apoferritin were used to observe particle

144

morphologies and to identify mineral phases. To identify the phases of Mn(III)

145

(oxy)hydroxide nucleated inside apoferritin, samples were first centrifuged at 14,800 rpm

146

for 10 minutes to separate potential homogeneous nucleation. The supernatants were

147

collected for ultracentrifuging (Thermo Scientific Sorvall WX Ultra Series Centrifuge with

148

a T-865 Fixed Angle Rotor) at 40,000 rpm for 30 minutes to obtain concentrated samples

149

of Mn-reconstituted ferritin. D-spacing data calculated from the HRTEM-electron

150

diffraction (ED) pattern were compared with the literature to further determine the phases

151

of reaction products.

152

In addition to ED analyses, wide angle X-ray scattering (WAXS) further confirmed

153

the mineral phase, formed inside apoferritin cavities. The measurements were conducted

154

with an energy of 58.290 keV (λ = 0.2127 Å) on the beamline 11ID-B at the Advanced

155

Photon Source at Argonne National Laboratory, IL. A sample reacted for 12 hrs at pH 9.0

156

with an Mn to apoferritin ratio of 2000:1 was transferred to a quartz capillary for in situ

7

ACS Paragon Plus Environment

Crystal Growth & Design

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

157

WAXS. The sample was exposed for 180 s. Data from a background sample, prepared

158

identically except for adding 200 µM of Mn2+(aq), was also obtained with the same beam

159

exposure time for background subtraction.

160

Particle size distributions for samples with and without apoferritin were measured

161

using dynamic light scattering (DLS, Zetasizer) at different elapsed times. Systems under

162

each pH condition were given enough time to react so that the shift in the nucleation

163

pathway from inside apoferritin cavities to in solution was completely observed in terms

164

of particle size change. The particle size evolutions were monitored for 12, 8, and 5 hr for

165

samples prepared at pH 8.9, 9.0, and 9.1, respectively. (S3 in the Supporting Information).

166

Colorimetric quantification of Mn(III) (oxy)hydroxide formation in apoferritin

167

Leucoberbelin blue (LBB, Sigma Aldrich) is a reducing agent that specifically

168

reduces Mn of higher oxidation states to Mn2+, forming a colored species that allows

169

colorimetric quantification of the oxidized Mn concentration using UV-vis spectroscopy.51

170

LBB solution was prepared by dissolving 0.004 % (w/v) of LBB in de-ionized water, to

171

which 45 mM acetic acid was added, and the solution was stored at 4 °C overnight.51 The

172

calibration curve was prepared by oxidizing the LBB solution with standard KMnO4

173

solutions, with the extinction coefficient, ɛ, determined to be 205,000 M-1 at a wavelength

174

of 625 nm.

175

To prepare the samples for colorimetric measurement, an aliquot of 0.3 mL was

176

taken from the reacting solutions and 1.5 mL of LBB solution was added. The samples

177

were allowed to equilibrate in the dark for at least 30 minutes, after which the absorbance

178

at 625 nm was measured using UV-Vis spectroscope (UV-Vis, Cary 50 Bio UV-Vis

179

spectroscopy, Varian Inc.,).

8

ACS Paragon Plus Environment

Page 10 of 34

Page 11 of 34

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

Crystal Growth & Design

180

To understand how LBB molecules gain access to Mn(III) mineral cores, we

181

analyzed the apoferritin particle size and morphology at varied pH conditions using atomic

182

force microscopy (AFM, Nanoscope V multimode SPM, Veeco Inc.) in tapping mode.

183

TEM images of samples after LBB treatment were also obtained to support the validity of

184

using the LBB method for quantifying heterogeneously nucleated Mn(III) inside the

185

apoferritin cavity. To visualize the protein, the TEM samples were negatively stained with

186

uranyl acetate (Electron Microscopy Sciences) and washed thoroughly with DI water.

187

Kinetic analyses

188

Equation (1) was hypothesized to depict the formation kinetics of -MnOOH,

189

which we identified to be the phase of the reaction product formed inside apoferritin (S4

190

in the Supporting Information).

191

= k[Mn2+]a [OH-]b

(1)

192

The LBB-probed -MnOOH concentrations over time were then fitted linearly.

193

The rate of Mn(III) formation at each experimental condition was then calculated from the

194

slope of the linear regression lines. Non-linear regression fittings then provided the best-

195

fit values of the overall rate constant for the combined oxidation and nucleation of

196

α-MnOOH and the orders of reaction with respect to concentrations of Mn2+ and OH-.

197

Based on the kinetic parameters obtained, we suggested the best predicted stoichiometric

198

numbers of reacting species (a and b in Equation (1)) and mechanism for the chemical

199

reactions involved. The proposed mechanism was further confirmed from the detection of

200

low concentrations of H2O2 by the peroxidase-catalyzed n,n-diethyl-p-phenylenediamine

201

oxidation method (S4 in the Supporting Information).

9

ACS Paragon Plus Environment

Crystal Growth & Design

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

202

Results and Discussion

203

Identification of -MnOOH nanoparticle formation inside the apoferritin cavities

204

Using XPS and TEM analyses, we identified two distinct nucleation pathways of

205

-MnOOH nanoparticle formation inside apoferritin cavities, as well as and larger-sized

206

Mn(OH)2 (s) particles (with an oxidized surface layer of Mn3O4) formed in bulk solution

207

not associated with apoferritin. The oxidation states of products formed from reactions in

208

the presence of apoferritin were measured using XPS (Figure 1a). The samples prepared

209

with apoferritin had a Mn 3p peak position at 48.6 eV, which is in accord with the average

210

of Mn(III) values reported in the literature and with our own reference samples (48.7 eV)

211

(Table S1).

212

To further characterize the phases of nucleated products inside apoferritin cavities,

213

TEM images together with electron diffraction (ED) patterns were obtained for the sample

214

prepared at pH 9 and with 200 µM Mn2+ (Figure 1b). Because Mn(OH)2 (s) forms more

215

readily with an elevated Mn2+ concentration in solution, the lowest Mn(II)/apoferritin ratio

216

(2000:1) was chosen for TEM analyses of nucleation inside apoferritin to avoid possible

217

confusion from Mn(OH)2 (s) formed in solution. D-spacings of the solid phase were

218

subsequently calculated from the ED pattern and compared with reference values of

219

manganite (-MnOOH), groutite (α-MnOOH), feitknechtite (-MnOOH), pyrochroite

220

(Mn(OH)2 (s)), and hausmannite (Mn3O4). Although both groutite and manganite have a

221

d-spacing value of around 2.67 Å, as observed by TEM (Figure 1b), ED analyses clearly

222

showed two large d-spacings of 5.14 and 4.45 Å, which match only with the reference

223

d-spacings of groutite (α-MnOOH) faces of (200) and (101). In addition to ED analyses,

224

from in situ measurements of synchrotron-based wide angle X-ray scattering (WAXS)

10

ACS Paragon Plus Environment

Page 12 of 34

Page 13 of 34

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

Crystal Growth & Design

225

(Figure 1c), we also clearly observed the strong diffraction pattern of α-MnOOH at 21°,

226

which is the strongest diffraction of α-MnOOH. The WAXS results further support the ED

227

analyses by confirming the phase of the nucleation product inside apoferritin cavities to be

228

α-MnOOH. Thus, we can rule out possible effects of ultracentrifugation and drying on the

229

phase transformation of synthesized α-MnOOH nanoparticles in apoferritin. Our finding

230

differs from that provided by Mann’s group, who found that the phase of ferritin-

231

encapsulated Mn resembles - and -MnOOH, based on X-ray absorption spectroscopy

232

(XAS).44 The difference in mineral phase formation between our study and Mann’s study

233

could arise because of varied experimental conditions. The concentration of apoferritin

234

(Mann’s study used 2.25 µM, while we used a much lower concentration, 0.10 µM) could

235

affect the phases of final reaction products due to different saturation conditions with

236

respect to the potential Mn(III) (oxy)hydroxide phases.44

237

To complement what we found for solid phase Mn(III) formation inside apoferritin

238

cavities, the phase of Mn (hydr)oxide particles formed in solution (from samples prepared

239

without apoferritin) were identified by XPS, XRD, and TEM (Figure 1a and Figure S5).

240

The XPS spectrum shows the peak position for a sample prepared without apoferritin is at

241

48.2 eV, which is in between the reference values of Mn(II) (48.1 eV), and Mn(III) (48.7

242

eV), indicating the formation of a mineral phase different from that found in the sample

243

prepared with apoferritin (i.e., a peak shown at 48.6 eV). Further characterization by XRD

244

and TEM confirms that the phase of nucleation in the bulk solution is mainly pyrochroite

245

(Mn(OH)2 (s)) with a small hint of oxidized hausmannite (Mn3O4) on the surface, with a

246

much bigger particle size (> 50 nm) (Figure S5b) than that of nanoparticles nucleated in

247

apoferritin (~8 nm) (Figure 1b). Comparison between nucleation in solution and that within

11

ACS Paragon Plus Environment

Crystal Growth & Design

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

248

apoferritin cavities suggests that apoferritin affects Mn oxidation and nucleation by altering

249

the phase of the reaction product and controlling the particle size.

250

To monitor the evolution of the dominant nucleation pathway in Mn-apoferritin

251

systems, from inside apoferritin cavities to in the bulk solution, we measured the particle

252

size distributions of samples prepared under varied experimental conditions (pH 8.9–9.1

253

and Mn(II)/protein ratios 2000:1–4000:1) at different elapsed times. Representative

254

particle size distributions under pH 8.9, 9.0, and 9.1 with Mn(II)/protein ratios 3500:1 are

255

shown in Figure 2 and the complete data set is available in Figure S3. The sizes of the

256

particles remained around 11 nm at the early stage of reaction (before 8, 6, and 4 hr,

257

respectively for samples prepared under pH 8.9, 9.0, and 9.1), indicating that the particles

258

formed mainly inside the apoferritin cavity, which has an outer diameter of 12 nm.19 As

259

the reaction continues, sharp increase in particle size suggests nucleation in solution

260

dominates over apoferritin in number concentration. Together with XPS and TEM

261

characterizations, the DLS data suggested there are two distinct nucleation pathways: one

262

inside the apoferritin and another in the bulk solution (Figure S3 and Figure S6). Mn

263

oxidation and mineralization inside apoferritin cavities forms α-MnOOH nanoparticles

264

with sizes confined to the dimension of the cavity (Figure S3). Nucleation in the bulk

265

solution forms combined pyrochroite (Mn(OH)2 (s)) and hausmannite (Mn3O4) with larger

266

sizes (> 50 nm) upon oxidation by dissolved oxygen (Figure S5 and Figure S6). The shift

267

in particle size profiles for samples prepared under varied experimental conditions also

268

indicate that nucleation inside apoferritin cavities is preferred to that in solution at the early

269

period of reaction (before 8, 6, and 4 hr respectively for samples prepared under pH 8.9,

270

9.0, and 9.1). The thermodynamics suggests that the preference could be due to the reduced

12

ACS Paragon Plus Environment

Page 14 of 34

Page 15 of 34

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

Crystal Growth & Design

271

energy barrier provided by iron-coordinating amino acid residues (His65, Glu27, Glu61,

272

Glu62, and Glu 107 for ferritins) at the ferroxidase center that catalyzes Mn oxidation,

273

similar to the reaction pathway in the protein catalysis mechanism reported in iron

274

deposition studies.52 As the nucleation sites in apoferritin approach saturation, nucleation

275

formed in bulk solution dominates as the reaction proceeds. We also found that the

276

transition between the two nucleation pathways is highly dependent on the initial

277

experimental conditions (Figure S3). The transition from nucleation inside apoferritin to

278

nucleation in bulk solution takes place earlier at higher pH and with higher Mn2+

279

concentration, implying that OH- and Mn2+ are rate-determining agents for apoferritin-

280

mediated heterogeneous oxidation and nucleation.

281

LBB quantification of -MnOOH nanoparticles formed inside apoferritin cavities

282

To obtain the combined rate of both oxidation and nucleation of Mn inside

283

apoferritin, α-MnOOH concentrations inside apoferritin cavities were quantified

284

colorimetrically by LBB, which is a reducing agent that specifically reduces Mn of higher

285

oxidation states to Mn2+. The LBB molecule, with its multiple benzene-ring structure, is

286

not likely to enter the protein’s interior through the hydrophilic channels because the

287

narrowest part of the channels is only 3 to 4 Å across.22 Previous research by Kim et al.

288

discovered that the apoferritin structure will undergo gradual disintegration under pH

289

below 3.40.47 Thus, we hypothesized that LBB penetrates the protein shell through the

290

disintegrated protein and hence reduces the α-MnOOH core. To test this hypothesis, the

291

particle size profiles of apoferritin solutions under pH 9.0, 3.0, and 1.0 were captured with

292

DLS. The large z-average (the intensity weighted mean hydrodynamic size) for pH 9.0

293

might result from the presence of impurities. Despite the interference from the background

13

ACS Paragon Plus Environment

Crystal Growth & Design

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

294

condition, as shown in Figure 3d, the z-average of the apoferritin samples increases

295

significantly with lower pH, which confirms that aggregation occurred as regular structures

296

of apoferritin were disrupted. AFM images confirmed the effects of pH on apoferritin

297

morphologies (Figure 3a–c). At pH 9.0, apoferritins generally retain the well-defined

298

spherical shape, suggesting intact structures. As the pH was lowered to 3.0, the particle

299

morphologies became less identifiable, and aggregations of disassembled apoferritin

300

fragments appear in the AFM images. At pH 1.0, the tertiary structures of the proteins were

301

completely lost, and an elongated conformation of unfolded polypeptides was observed.

302

The pH of LBB solution used in our study was 3.1. Based on the AFM and DLS results,

303

we concluded that under the low pH condition induced by the LBB solution, LBB

304

molecules are able to traverse the protein shell and reduce the α-MnOOH core in the central

305

cavity through fragmented surfaces resulting from the acidic pH environment.

306

To further confirm the successful reduction of the α-MnOOH by the LBB method,

307

TEM images of samples after LBB treatment were obtained (Figure 3e). The negative

308

staining shows protein aggregation in post-LBB treated samples, again confirming the

309

deterioration of apoferritin in the LBB solution. Also, we detected d-spacing values of 3.30

310

Å, which are consistent with the reference values of uranyl acetate (Figure 3e), while no

311

values that could possibly match with Mn (hydr)oxide phases were identified. The absence

312

of solid phase Mn(III) in TEM images (Figure 3e) suggests that the α-MnOOH particles

313

formed inside apoferritin cavity have been completely reduced to Mn2+, and thus the LBB

314

method provides a reliable means for quantifying α-MnOOH deposition inside apoferritin

315

cavities.

14

ACS Paragon Plus Environment

Page 16 of 34

Page 17 of 34

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

Crystal Growth & Design

316

Determination of the combined rates of oxidation and nucleation of Mn inside

317

apoferritin from pseudo-zeroth order kinetics

318

To determine the combined rate of Mn oxidation and nucleation inside apoferritin using

319

the LBB method, we measured the concentrations of Mn(III) in samples prepared under all

320

experimental conditions at different elapsed times. The combined oxidation and nucleation

321

process inside apoferritin obeyed pseudo-zeroth order kinetics, with Mn2+ and OH- being

322

the rate-determining agents. As shown in Figure S3, particle size profiles of the samples

323

indicated that the transition of nucleation from inside apoferritin to in the bulk solution

324

depends on both pH and Mn2+ concentrations. Therefore, to obtain accurate kinetics solely

325

for nucleation occurring inside apoferritin cavities, we restricted the duration of the

326

reaction to before the transition, so that effect of Mn(OH)2 (s) formation in the bulk solution

327

was not taken into account. For pH 8.9 and 9.0, concentrations of Mn(III) were measured

328

at 2 hr intervals to 8 hr; for pH 9.1, concentrations of Mn(III) were measured at 0.5 hr

329

intervals to 3 hr because of the markedly more rapid MnOOH formation from

330

heterogeneous nucleation at the slightly higher pH. The saturation indices of Mn(OH)2 (s)

331

were calculated to explain the more rapid nucleation in solution with increasing pH and

332

Mn2+ concentrations (Table S2 and S3 in the Supporting Information). α-MnOOH

333

concentration profiles for each experimental condition are shown in Figure 4. A linear

334

increase in Mn(III) concentrations with time is seen and can be explained from the pseudo-

335

zeroth order kinetics. Because unlimited access to air was guaranteed for all samples during

336

reaction, and because oxygen consumption by Mn oxidation was negligible compared to

337

the total dissolved oxygen in the solution, the oxygen level was assumed to remain

338

unchanged throughout the reaction. Also, the Mn concentration was assumed to be constant

15

ACS Paragon Plus Environment

Crystal Growth & Design

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

339

due to the high Mn atom to apoferritin ratio (higher than 1000:1) used in this study.

340

Therefore, we fitted the experimental data with linear regression lines and obtained the

341

initial rate of reaction by calculating the slope of each (Figure 4). The initial rate of reaction

342

increases with pH and Mn2+ concentrations, which confirms that the OH- and Mn2+ are

343

rate-determining agents.

344

Overall reaction rates and the proposed mechanism

345

By fitting the overall rates of oxidation and nucleation obtained from experiments

346

to the hypothesized rate equation, Equation (1), we specified the kinetic parameters in the

347

rate equation and proposed that -MnOOH nanoparticles form in sequential steps of

348

oxidation and mineralization at the ferroxidase center of apoferritin with the involvement

349

of two intermediates. To obtain the rate constant, k, and the orders of reactions with regard

350

to OH- and Mn2+, a and b as specified in Equation (1), we fitted the experimentally obtained

351

rates of reaction to Equation (1) with assumed combinations of a and b, using non-linear

352

least squares regression (Figure 5). The rate equation best describes the experimental data

353

when a and b are 2 and 4, respectively, under which case the best-fit value of the overall

354

rate constant, k, is 1.28  10-9 M-5hr-1 (Figure 5). The high order of reaction with respect

355

to OH- also explains the high sensitivity of the reaction systems to even minor changes in

356

pH condition (Figure 5b). To further support the validity of the fitting results, we analyzed

357

the goodness of fit by calculating R squares for a range of a and b values. The best-fit

358

results give a R square value of 0.9953, confirming the high accuracy with which the model

359

describes the experimental data (S7 in the Supporting Information).

360 361

Based on the kinetics parameters predicted by the computational fitting, we propose here a possible mechanism for enzymatically-catalyzed Mn(III)OOH formation.

16

ACS Paragon Plus Environment

Page 18 of 34

Page 19 of 34

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

Crystal Growth & Design

362

Oxidation: 2Mn2+ + O2 + 4OH- + P

P-[MnIII2O2(OH)2] + H2O2

(2)

363

Disassociation: P-[MnIII2O2(OH)2]

[MnIII2O2(OH)2] + P

(3)

364

Mineralization: [MnIII2O2(OH)2]

365

Overall: 2Mn2+ + O2+ 4OH- → 2MnIIIOOH(core) + H2O2

2MnIIIOOH(core)

(4) (5)

366

Here, P is the protein, k1, k2, and k3 are the rate constants for the respective

367

elementary steps. We propose that sequential oxidation and mineralization reactions with

368

the formation of two intermediates lead to the overall α-MnOOH deposition in apoferritin.

369

The rate-determining step in the oxidation process generates the Mn(III)-apoferritin

370

complex, which forms at the ferroxidase dinuclear center because of protein catalysis. The

371

oxidized Mn(III) then rapidly dissociate from the binding sites in the protein shell. The free

372

Mn(III) subsequently diffuses into the protein cavities, where mineralization occurs to form

373

nascent -MnOOH core. The formation of intermediates in the Mn-apo system can be

374

justified from the iron deposition process in native ferritin. Previous studies reported that

375

the first intermediate formed from Fe(II) oxidation at the ferroxidase center by oxygen is

376

μ-1,2-peroxodiFe(III).26, 53-56 The initial intermediate subsequently dissociates to one or

377

more μ-oxo(hydroxo)-bridged diFe(III) intermediate(s), which ultimately lead to the

378

formation of the nascent mineral core. Based on the Fe studies, enzymatically catalyzed

379

oxidation reactions occurring at the ferroxidase dinuclear center might also exist for the

380

Mn-apoferritin system and explain the observed kinetics in this study, although the phase

381

and structure of the intermediates formed in the two systems differ from each other because

382

of the difference in the reaction stoichiometry.

383 384

In this study, by assuming pseudo-steady state kinetics for P-[Mn2O2(OH)2], we derived the following expression for P-[Mn2O2(OH)2]: 17

ACS Paragon Plus Environment

Crystal Growth & Design

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

385

Page 20 of 34

P-[Mn2O2(OH)2] =k1/k2[O2][Mn2+]2[OH-]4.

(6)

386

Next, by assuming pseudo-steady state kinetics for [Mn2O2(OH)2] and substitute Equation

387

(6) into the expression, we obtained Equation (7) for [Mn2O2(OH)2]

388

[Mn2O2(OH)2] = k2/k3 P-[Mn2O2(OH)2] = k1/k3[O2][Mn2+]2[OH-]4.

389

(7)

390

By substituting Equation (7) into the rate equation for α-MnOOH formation (Equation (1)),

391

we obtained Equation (8)

392



= k1[O2][Mn2+]2[OH-]4.

(8)

393

The product of the elementary rate constant, k1 and dissolved oxygen concentration can be

394

lumped into the overall rate constant, k, and the rate equation can be rearranged into the

395

final form

396

= k[Mn2+]2[OH-]4.

(9)

397

Because the overall reaction forms H2O2 as one of the final products (Equation (4)),

398

to further test the validity of the proposed reaction stoichiometry, we used the peroxidase-

399

catalyzed n,n-diethyl-p-phenylenediamine oxidation method to detect the presence of H2O2

400

(S4 in the Supporting Information). The UV spectra obtained after 4 hr and 12 hr of reaction

401

show clear absorbance peaks at 551 nm.57 Because of the instability of H2O2, we focused

402

on confirming the presence of H2O2 rather than obtaining the exact H2O2 concentration.

403

The presence of strong peak at 551 nm (Figure S4) confirmed the formation of H2O2 during

18

ACS Paragon Plus Environment

Page 21 of 34

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

Crystal Growth & Design

404

the overall reaction. Based on the intermediate formation and the presence of H2O2, we

405

suggest that separate oxidation and mineralization reactions are involved in the overall

406

MnOOH formation in apoferritin.

407

To extend our discussion of the reaction stoichiometry and the catalytic function of

408

apoferritin on α-MnOOH formation to the synthetic processes of other nanomaterials, we

409

compared the deposition of Mn in apoferritin to that of Fe. As shown in Equation (5), the

410

overall stoichiometry involved in α-MnOOH formation is similar to the protein catalysis

411

mechanism proposed in Fe studies in the respect that both systems undergo apoferritin-

412

mediated oxidation reactions.29, 58 The similarity suggests that Mn2+ up-taken into the

413

protein is likely to go through the ferroxidase center, where a series of key glutamate

414

residues on the exposed surface serve to catalyze the oxidation of Mn.20, 29, 58 However, the

415

deposition of Mn in apoferritin is distinguished from Fe(III) oxy(hydr)oxide formation in

416

that the order of reactions with respect to the metal and hydroxyl are 2 and 2, respectively

417

in the protein catalysis model proposed in the Fe study, while we found the orders were 2

418

and 4 for the Mn-apoferritin system. This difference in kinetic aspect suggests that despite

419

protein catalyzes the oxidation reactions in both cases, the reactions themselves are

420

different, forming intermediates of different phases and structures. In addition, the protein

421

catalysis mechanism is applicable for even high Mn(II)/apoferritin ratios (higher than 1000

422

:1). In the Fe case, the mechanism occurs only transiently at the incipient stage of reaction

423

under low Fe flux (less than 48 Fe(II)/protein), and reactions occur directly on the mineral

424

surface without protein catalysis as Fe flux is increased above the threshold ratio.29, 58 The

425

slower oxidation kinetics of Mn might have contributed to this difference in enzymatic

426

function of apoferritin in Mn and Fe systems. Despite the difference in phases of

19

ACS Paragon Plus Environment

Crystal Growth & Design

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

427

intermediates and kinetics, this mechanistic study also elucidates the role of the ferroxidase

428

center shared between the two systems, suggesting that the enzymatic sites of the protein

429

may engage in the formation of other metal-oxyhydroxide(s) in similar ways. Thus, this

430

work highlights the importance of kinetic control in mechanistic studies or nanomaterial

431

synthesis when using apoferritin as the mineralization platform. This new information can

432

benefit in the optimization of bio-inspired synthesis of uniform-sized nanomaterials.

433

Conclusions

434

We investigated the kinetics and mechanism of Mn oxidation and nucleation in

435

apoferritin. We identified two distinct nucleation pathways in an apoferritin-Mn system:

436

nucleation inside apoferritin cavities forms α-MnOOH nanoparticles with sizes of 6–8 nm,

437

and nucleation in solution forms Mn(OH)2 (s) (with an oxidized layer of Mn3O4) particles

438

of larger sizes (> 50 nm). Nucleation of α-MnOOH in apoferritin is favored at the early

439

stage of reaction, then nucleation of Mn (hydr)oxides in solution gradually dominates as

440

the reaction continues. We confirmed that the step-wise disassembly of apoferritin under

441

acidic conditions induced by LBB solution allows the LBB molecules to traverse the

442

protein shell, and thus quantify the nucleated α-MnOOH in apoferritin as a complete

443

reduction of Mn(III). From the fitting results, we found that Mn2+ and OH- are rate-

444

determining agents with best-fit orders of reaction of 2 and 4, respectively. Based on kinetic

445

analyses, we proposed that α-MnOOH forms along with H2O2 through sequential steps of

446

oxidation and mineralization, with the possible involvement of two intermediates. Our

447

findings illustrate the appropriate experimental conditions that allow pure synthesis of α-

448

MnOOH nanoparticles in apoferritin without being affected by unwanted Mn (hydr)oxide

449

nucleation in solution. In addition, the suggested oxidation and nucleation mechanisms 20

ACS Paragon Plus Environment

Page 22 of 34

Page 23 of 34

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

Crystal Growth & Design

450

will provide implications to understand the formation behaviors of other inorganic

451

nanoparticles using other protein cages as macromolecular templates through

452

biomineralization.

453

Associated Contents

454

Supporting Information. Dissolved oxygen (S1), XPS references (S2), particle size

455

profiles for samples with 0.1 M apoferritin solution and saturation indices (S3), N,N-

456

diethyl-p-phenylenediamine (DPD) method for detection of H2O2 (S4), phase

457

identification of nucleation of Mn (hydro)oxides in solution (S5), particle size profiles for

458

samples without apoferritin solution (S6), and evaluation of fitting results (S7). The

459

Supporting Information is available free of charge on the ACS Publications website.

460

Author Information

461

Corresponding Author

462

*E-mail: [email protected]

463

Author Contributions

464

The manuscript was written through contributions of all authors. All authors have given

465

approval to the final version of the manuscript.

466



467

Notes

468

The authors declare no competing financial interests.

469

Acknowledgments

Y. H. and H. J. contributed equally.

470

The authors would like to acknowledge the support from the National Science

471

Foundation’s Environmental Chemical Sciences program (CHE-1610728). We also thank 21

ACS Paragon Plus Environment

Crystal Growth & Design

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

472

Washington University’s Institute of Materials Science & Engineering (IMSE) for use of

473

XPS and TEM, and Professor James C. Ballard for carefully reviewing the manuscript.

474

Work at the Advanced Photon Source (Sector 11 ID-B) at Argonne National Laboratory

475

was supported by the US Department of Energy, Office of Science, Office of Basic Energy

476

Sciences, under Contract No. DE-AC02-06CH11357.

22

ACS Paragon Plus Environment

Page 24 of 34

Page 25 of 34

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

Crystal Growth & Design

References (1) Kashyap, S.; Woehl, T. J.; Liu, X.; Mallapragada, S. K.; Prozorov, T., ACS Nano 2014, 8, 9097-9106. (2) Theil, E. C., Curr. Opin. Chem. Biol. 2011, 15, 304-311. (3) Bode, S. A.; Minten, I. J.; Nolte, R. J.; Cornelissen, J. J., Nanoscale 2011, 3, 2376-2389. (4) Flenniken, M. L.; Willits, D. A.; Brumfield, S.; Young, M. J.; Douglas, T., Nano Lett. 2003, 3, 1573-1576. (5) Douglas, T.; Strable, E.; Willits, D.; Aitouchen, A.; Libera, M.; Young, M., Adv. Mater. 2002, 14, 415-418. (6) Douglas, T.; Young, M., Nature 1998, 393, 152-155. (7) Mann, S.; Meldrum, F. C., Adv. Mater. 1991, 3, 316-318. (8) Jutz, G. n.; van Rijn, P.; Santos Miranda, B.; Böker, A., Chem. Rev. 2015, 115, 16531701. (9) Mougin, N. C.; van Rijn, P.; Park, H.; Müller, A. H. E.; Böker, A., Adv. Funct. Mater. 2011, 21, 2470-2476. (10) Lin, X.; Xie, J.; Niu, G.; Zhang, F.; Gao, H.; Yang, M.; Quan, Q.; Aronova, M. A.; Zhang, G.; Lee, S.; Leapman, R.; Chen, X., Nano Lett. 2011, 11, 814-819. (11) Uchida, M.; Kang, S.; Reichhardt, C.; Harlen, K.; Douglas, T., Biochim. Biophys. Acta. 2010, 1800, 834-845. (12) Li, M.; Viravaidya, C.; Mann, S., Small 2007, 3, 1477-1481. (13) Douglas, T.; Dickson, D. P. E.; Betteridge, S.; Charnock, J.; Garner, C. D.; Mann, S., Science 1995, 269, 54-57. (14) Allen, M.; Willits, D.; Young, M.; Douglas, T., Inorg. Chem. 2003, 42, 6300-6305. (15) Uchida, M.; Klem, M. T.; Allen, M.; Suci, P.; Flenniken, M.; Gillitzer, E.; Varpness, Z.; Liepold, L. O.; Young, M.; Douglas, T., Adv. Mater. 2007, 19, 1025-1042. (16) Okuda, M.; Iwahori, K.; Yamashita, I.; Yoshimura, H., Biotechnol. Bioeng. 2003, 84, 187-194. (17) Polanams, J.; Ray, A. D.; Watt, R. K., Inorg. Chem. 2005, 44, 3203-3209. (18) de la Escosura, A.; Verwegen, M.; Sikkema, F. D.; Comellas-Aragones, M.; Kirilyuk, A.; Rasing, T.; Nolte, R. J.; Cornelissen, J. J., Chem. Commun. 2008, 1542-1544. (19) Ueno, T.; Suzuki, M.; Goto, T.; Matsumoto, T.; Nagayama, K.; Watanabe, Y., Angew. Chem. 2004, 116, 2581-2584. (20) Chasteen, N. D.; Harrison, P. M., J. Struct. Biol. 1999, 126, 182-194. (21) Harrison, P. M.; Arosio, P., Biochim. Biophys. Acta. 1996, 1275, 161-203. (22) Bakker, G. R.; Boy, R. F., J. Biol. Chem. 1986, 261, 13182-13185. (23) Michel, F. M.; Hosein, H. A.; Hausner, D. B.; Debnath, S.; Parise, J. B.; Strongin, D. R., Biochim. Biophys. Acta. 2010, 1800, 871-885. (24) Pereira, A. S.; Tavares, P.; Lloyd, S. G.; Danger, D.; Edmondson, D. E.; Theil, E. C.; Huynh, B. H., Biochemistry 1997, 36, 7917-7927. (25) Sun, S.; Arosio, P.; Levi, S.; Chasteen, N. D., Biochemistry 1993, 32, 9362-9369. (26) Bou-Abdallah, F.; Zhao, G.; Mayne, H. R.; Arosio, P.; Chasteen, N. D., J. Am. Chem. Soc 2005, 127, 3885-3893. (27) Zhao, G.; Bou-Abdallah, F.; Arosio, P.; Levi, S.; Janus-Chandler, C.; Chasteen, N. D., Biochemistry 2003, 42, 3142-3150. (28) Jameson, G. N. L.; Jin, W.; Krebs, C.; Perreira, A. S.; Tavares, P.; Liu, X.; Theil, E. C.; Huynh, B. H., Biochemistry 2002, 41, 13435-13443. 23

ACS Paragon Plus Environment

Crystal Growth & Design

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

(29) Pereira, A. S.; Small, W.; Krebs, C.; Tavares, P.; Edmondson, D. E.; Theil, E. C.; Huynh, B. H., Biochemistry 1998, 37, 9871-9876. (30) Douglas, T.; Stark, V. T., Inorg. Chem. 2000, 39, 1828-1830. (31) Okuda, M.; Kobayashi, Y.; Suzuki, K.; Sonoda, K.; Kondoh, T.; Wagawa, A.; Kondo, A.; Yoshimura, H., Nano Lett. 2005, 5, 991-993. (32) Yamashita, I.; Hayashi, J.; Hara, M., Chem. Lett. 2004, 33, 1158-1159. (33) Warne, B.; Kasyutich, O. I.; Mayes, E. L.; Wiggins, J. A.; Wong, K. K., IEEE Trans. Magn. 2000, 36, 3009-3011. (34) MaHam, A.; Tang, Z.; Wu, H.; Wang, J.; Lin, Y., Small 2009, 5, 1706-1721. (35) Lei, Y.; Hamada, Y.; Li, J.; Cong, L.; Wang, N.; Li, Y.; Zheng, W.; Jiang, X., J. Control. Release 2016, 232, 131-142. (36) Truffi, M.; Fiandra, L.; Sorrentino, L.; Monieri, M.; Corsi, F.; Mazzucchelli, S., Pharmacol. Res. 2016, 107, 57-65. (37) Liang, M.; Fan, K.; Zhou, M.; Duan, D.; Zheng, J.; Yang, D.; Feng, J.; Yan, X., Proc. Nat. Acad. Sci. 2014, 111, 14900-14905. (38) Schwarz, B.; Douglas, T., Wiley Interdiscip. Rev.: Nanomed. Nanobiotechnol. 2015, 7, 722-735. (39) Patterson, D. P.; Rynda-Apple, A.; Harmsen, A. L.; Harmsen, A. G.; Douglas, T., ACS nano 2013, 7, 3036-3044. (40) Ferenc, K.; Geninatti Crich, S.; Aime, S., Angew. Chem. Int. Ed 2010, 49, 612-615. (41) Sana, B.; Poh, C. L.; Lim, S., Chem. Commun. 2012, 48, 862-864. (42) Terreno, E.; Dastru, W.; Delli Castelli, D.; Gianolio, E.; Geninatti Crich, S.; Longo, D.; Aime, S., Curr. Med. Chem. 2010, 17, 3684-3700. (43) Meldrum, F. C.; Douglas, T.; Levi, S.; Arosio, P.; Mann, S., J. Inorg. Biochem. 1995, 58, 59-68. (44) Mackle, P.; Charnock, J. M.; Garner, C. D.; Meldrum, F. C.; Mann, S., J. Am. Chem. Soc 1993, 115, 8471-8472. (45) Arosio, P.; Adelman, T. G.; Drysdale, J. W., J. Biol. Chem. 1978, 253, 4451-4458. (46) Levi, S.; Salfeld, J.; Franceschinelli, F.; Cozzi, A.; Dorner, M. H.; Arosio, P., Biochemistry 1989, 28, 5179-5184. (47) Kim, M.; Rho, Y.; Jin, K. S.; Ahn, B.; Jung, S.; Kim, H.; Ree, M., Biomacromolecules 2011, 12, 1629-1640. (48) Jung, H.; Jun, Y.-S., Environ. Sci. Technol. 2016, 50, 105-113. (49) Jung, H.; Lee, B.; Jun, Y.-S., Langmuir 2016, 32, 10735-10743. (50) Ilton, E. S.; Post, J. E.; Heaney, P. J.; Ling, F. T.; Kerisit, S. N., Appl. Surf. Sci. 2016, 366, 475-485. (51) Tebo, B. M.; Clement, B. G.; Dick, G. J.; Hurst, C.; Crawford, R.; Garland, J.; Lipson, D.; Mills, A.; Stetzenbach, L., Man. Environ. Microbiol. 2007, 1223-1238. (52) De Yoreo, J. J.; Vekilov, P. G., Rev. Mineral. Geochem. 2003, 54, 57-93. (53) Bou-Abdallsh, F.; Papaefthymiou, G. C.; Scheswohl, D. M.; Sanga, S. D.; Arosio, P.; Chasteen, N. D., Biochem. J. 2002, 364, 57-63. (54) Bauminger, E. R.; Harrison, P. M.; Hechel, D.; Nowik, I.; Treffry, A., Biochim. Biophys. Acta. 1991, 1118, 48-58. (55) Bauminger, E.; Harrison, P.; Hechel, D.; Hodson, N.; Nowik, I.; Treffry, A.; Yewdall, S., Biochem. J. 1993, 296, 709-719. (56) Hwang, J.; Krebs, C.; Huynh, B. H.; Edmondson, D. E.; Theil, E. C.; Penner-Hahn, J. E., Science 2000, 287, 122-125. 24

ACS Paragon Plus Environment

Page 26 of 34

Page 27 of 34

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

Crystal Growth & Design

(57) Bader, H.; Sturzenegger, V.; Hoigné, J., Water Res. 1988, 22, 1109-1115. (58) Yang, X.; Chen-Barrett, Y.; Arosio, P.; Chasteen, N. D., Biochemistry 1998, 37, 97439750.

25

ACS Paragon Plus Environment

Crystal Growth & Design

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

List of Figures Figure 1. (a) XPS spectra for samples prepared at pH 9.0 and with 200 µM Mn2+, with and without apoferritin. (b) TEM image and ED pattern of nucleated Mn(III) inside apoferritin prepared at pH 9.0 and with 200 µM Mn2+. (c) The background subtracted results of the in situ measurement of the synchrotron-based wide angle X-ray scattering used in identifying the phase of solid-state Mn(III) formation inside apoferritin cavities Figure 2. DLS particle size profiles for samples prepared under (a) pH 8.9, (b) pH 9.0, and (c) pH 9.1 with 350 M Mn2+. Figure 3.

AFM images of 0.1M apoferritin solution prepared under (a) pH 9, (b) pH 3,

and (c) pH 1. Height profiles were obtained at white-dotted lines. (d) Z-averages (intensity weighted mean hydrodynamic sizes) from DLS measurement for 0.1M apoferritin solution prepared under pH 9.0, 3.0, and 1.0. (e) TEM image of post-LBB treated sample prepared under pH 9.0 and with 200 M Mn2+, after negative staining with uranyl acetate. Figure 4. Experimentally measured concentrations of Mn(III) determined using the LBB method, and fitted linear regression lines for samples with different Mn(II)/protein ratios at (a) pH 8.9, (b) pH 9.0, and (c) pH 9.1. Figure 5.

Rates of overall Mn oxidation and nucleation inside apoferritin cavities

determined from the slopes of linear regression lines in Fig. 3 and fitted results using Equation (1), with a = 2, b = 4, and k = 1.28  10-9 M-5hr-1 with respect to (a) Mn2+ concentration and (b) pH.

26

ACS Paragon Plus Environment

Page 28 of 34

Page 29 of 34

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

Crystal Growth & Design

Figure 1

27

ACS Paragon Plus Environment

Crystal Growth & Design

3 hr 6 hr 8 hr 12 hr

40 30

Hetero. 20 Nucleation (< 8hr)

Homo. Nucleation (> 8hr)

10 0

1

10

100

Particle Size (nm)

1000

(c)

pH 9.0 2 hr 4 hr 6 hr 8 hr 10 hr

40 30 20

Homo. Nucleation (> 8hr)

Hetero. Nucleation (< 8hr)

10 0

1

10

100

Number Percentage (%)

(b)

pH 8.9

Number Percentage (%)

(a) Number Percentage (%)

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 30 of 34

1000

Particle Size (nm)

Figure 2

28

ACS Paragon Plus Environment

pH 9.1 1 hr 2 hr 3 hr 4 hr 5 hr

40 30 20

Homo. Nucleation (> 4hr)

Hetero. Nucleation (< 4hr)

10 0

1

10

100

Particle Size (nm)

1000

Page 31 of 34

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

Crystal Growth & Design

Figure 3

29

ACS Paragon Plus Environment

Crystal Growth & Design

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 4

30

ACS Paragon Plus Environment

Page 32 of 34

(a)

6

pH 8.9 pH 9.0 pH 9.1

4 2 0

200

Mn

(b)

Rate of Reaction (M/hr)

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

Crystal Growth & Design

Reaction rate (M/hr)

Page 33 of 34

6 4

2+

300

400 Concentration (M)

2000:1 2500:1 3000:1 3500:1 4000:1

2 0

8.9

9.0

pH

Figure 5

31

ACS Paragon Plus Environment

9.1

Crystal Growth & Design

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 of Contents Graphic and Synopsis (For Table of Contents Use Only)

Synopsis: This work investigated the kinetics of -MnOOH nanoparticle formation inside apoferritin cavities and proposed mechanisms for the Mn oxidation and nucleation reactions. Title: Kinetics of α-MnOOH Nanoparticle Formation through Enzymatically-catalyzed Biomineralization inside Apoferritin. Authors: Yue Hui†, Haesung Jung†, Kim Doyoon, and Young-Shin Jun*. †

These authors contributed equally *To whom correspondence should be addressed

32

ACS Paragon Plus Environment

Page 34 of 34