1172109

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

BNL-107589-2015-JAAM

BNL-107589-2015-JA

Quantitative and Qualitative Determination of Polysulfide Species in the


Electrolyte of Lithium-Sulfur Battery by HPLC ESI/MS with One-step
Derivatization

D. Zheng

To be published in "Advanced Energy Materials"

March 2015

Chemistry Department
Brookhaven National Laboratory

U.S. Department of Energy


USDOE Office of Energy Efficiency and Renewable Energy (EERE)

Notice: This manuscript has been authored by employees of Brookhaven Science Associates, LLC under
Contract No. DE-SC0012704 with the U.S. Department of Energy. The publisher by accepting the
manuscript for publication acknowledges that the United States Government retains a non-exclusive, paid-
up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow
others to do so, for United States Government purposes.
DISCLAIMER

This report was prepared as an account of work sponsored by an agency of the


United States Government. Neither the United States Government nor any
agency thereof, nor any of their employees, nor any of their contractors,
subcontractors, or their employees, makes any warranty, express or implied, or
assumes any legal liability or responsibility for the accuracy, completeness, or any
third party’s use or the results of such use of any information, apparatus, product,
or process disclosed, or represents that its use would not infringe privately owned
rights. Reference herein to any specific commercial product, process, or service
by trade name, trademark, manufacturer, or otherwise, does not necessarily
constitute or imply its endorsement, recommendation, or favoring by the United
States Government or any agency thereof or its contractors or subcontractors.
The views and opinions of authors expressed herein do not necessarily state or
reflect those of the United States Government or any agency thereof.
BNL-107589-2015-JA

DOI: 10.1002/((please add manuscript number))


Article type: Communication

Quantitative and Qualitative Determination of Polysulfide Species in the Electrolyte of


Lithium-Sulfur Battery by HPLC ESI/MS with One-step Derivatization

Dong Zheng, Deyu Qu, Xiao-Qing Yang, Xiqian Yu, Hung-Sui Lee and Deyang Qu*

Dr. Dong Zheng, Prof. Deyang Qu


Department of Chemistry, University of Massachusetts Boston, Boston, MA 02125, USA
E-mail: Deyang.qu@umb.edu
Prof. Deyu Qu
Department of Chemistry, School of Science, Wuhan University of Technology, Wuhan 430070,
Hubei, P.R. China
Dr. Xiao-Qing Yang, Dr. Xiqian Yu, Dr. Hung-Sui Lee
Chemistry Department, Brookhaven National Laboratory, Upton, NY 11973, USA
Keywords: polysulfide, Li-S battery, HPLC ESI/MS

A Li-S system promises extraordinary theoretical energy density (2600WhKg-1) [1]. Like many

battery systems, this impressive capacity is accompanied with a trade-off in complexity. In the case

of rechargeable Li-S batteries, such complexities result from the redox reaction of sulfur, which is

considered as one of the most complex electrochemical reactions involving the formation of

multiple polysulfide ions from multi-step, electron-transfer, homogeneous and heterogeneous

chemical reactions.

The polysulfide ions (Sn2- ) are generally soluble in non-aqueous electrolytes, the linear

longer chain Sn2- ions tend to diffuse away from the S electrode. The soluble polysulfide ions can

react with Li anode and form nonconductive deposits. The back and forth of polysulfide ions being

reduced at the negative electrode and then oxidized at the positive electrode is the root cause of the

so-called “shuttle effect” ; this effect can greatly accelerate the self-discharge and loss of

cyclability. In order to materialize the full potential of Li-S chemistry, the impacts of dissolved

polysulfide ions have to be alleviated [2,3]. Unfortunately, the detailed mechanisms for the sulfur

redox reaction in non-aqueous electrolytes are still not well understood, but it is consensus that the

dissolved polysulfide ions play critical roles in the reaction. To make things more complex, the

1
solubility and thermodynamic equilibrium of the polysulfide ions with various length are solvent

dependent [4-6]. However, even with decades of genuine efforts, there still is a lack of a reliable

analytical method to qualitatively separate and analyze the polysulfide ions of various lengths,

which hinder the comprehensive understand of the reaction mechanism.

Initially, UV-Visible spectroscopy was used as the major instrument to investigate the

soluble polysulfide ions in aprotic solvents, since polysulfide solutions are generally colored. It was

believed that the polysulfides of different types give different color [7-9]. Electro-analytical methods,

specifically cyclic voltammetry, were also used in an attempt to identify the polysulfide ions [10].

The mechanisms for the sulfur redox reactions as well as the chemical equilibriums among the

polysulfide ions were also studied by means of spectro-electrochemical methods [11-13]. However the

results from UV-Vis spectrometry and cyclic voltammetry provided limited information about the

polysulfide species due to the overlap of absorbance and redox peaks for different polysulfide

species, respectively. For example, broad and overlap absorption bands were demonstrated in UV-

Vis spectra, first derivative of the UV-Vis spectra was used to identify the number and position of

the bands [9]. Therefore, isolation of a population of single polysulfide chain lengths was proven to

be a challenge using UV-Vis [14].

Most recently, in-situ Raman spectroscopy was used to investigate the polysulfides formed

during the operation of a Li-S cell [15]. However, most of the Raman data for polysulfide were

obtained from crystalline solids, which is not helpful to interpret the Raman spectra of solvated

polysulfide ions. Therefore, theoretically calculated Raman spectra have to be used to assign the

polysulfide ions to the experimental peaks. Abruna et al reported on the use of X-ray absorption

near edge spectroscopy (XANES) to probe the polysulfide, however only the average oxidation

state of sulfur could be estimated [16].

The main issue for the instrumental analysis of polysulfides is that it is almost impossible to

prepare polysulfide anions with a precisely defined chain length in solution for reference

measurements, since disproportionation and redox reactions result in distribution of polysulfide


2
anions with various chain length [15]. Due to the different chain length and molar mass of different

polysulfide species, the powerful and widely used analytical method, chromatography tandem with

mass spectroscopy (MS), could be potentially used to analyze the polysulfide species, because no

reference of pure polysulfides are needed for the quantitative and qualitative measurement. The

high performance liquid chromatography (HPLC) tandem with electrospray-mass spectroscopy

(ESI-MS) was indeed introduced to analyze polysulfide specie [17]. By utilizing reverse phase (RP)

HPLC and ESI-MS, the polysulfide species formed during the charge and discharge were separated

and qualitatively determined. However, the analysis of polysulfide species by RP-HPLC/ESI-MS

without a specially developed assay is questionable [18]. Since reverse phase (RP) chromatography

is a typical partition chromatography, in which the stationary phase is nonpolar and mobile phase is

relatively polar. The retention mechanism for partition chromatography is based on the

intermolecular interaction among the mobile phase, analytes and stationary phase of the column. It’s

a method for the separation of covalent compounds and not suitable for the separation of ionic

compounds except for the macro-molecules of proteins and peptides. Polysulfide anions are

particularly difficult to be separated, even with electrophoretic and ion paring chromatography
[17,18]
, which are designed for separation of ionic compounds.

Lev et al reported on the analysis of polysulfide species in aqueous solution by means of gas

chromatography (GC) electron impact (EI)/MS with one step derivatization of the polysulfides [21].

The ionic polysulfide species were derivatized to form the corresponding covalent organic sulfur

species as shown in equation 1, which could then be separated and identified by GC EI/MS.

However, the derivatized polysulfide species, especially the ones with longer chain lengths will

decompose during the temperature ramp of a GC system; for example, (CH3)2Sn with n ≥6 cannot be

detected. Interestingly, it was shown that the derivatized polysulfide species can also be totally

resolved on reversed-phase HPLC system [21, 22]. Although the separation of derivatized species can

3
be achieved by a HPLC method, the identification of individual derivatized species in the

chromatogram with UV detector is still a challenge due to the lack of adequate polysulfide

standards, especially for the linear long chain length species. ESI-MS is a powerful tool for the

identification and qualification of derivatized species. But the lack of polar function group in the

methylated polysulfide species makes the ionization efficiency for those species low, which results

in the sensitivity of ESI-MS for those species to be low as shown in the supporting information

(Figure S-1). In this work, a novel derivatization reagent was reported and consequently the

derivatized polysulfide species were effectively analyzed by HPLC (+)ESI/MS. For comparison ,

the (-) ESI/MS through direct infusion and tandem with HPLC were also studied in this work.

Polysulfide species in an aprotic solution can be analyzed by directly introducing into the

ESI/MS. Figure 1A shows the (-) ESI/MS spectra of solution B through direct injection. Figure 1A

clearly demonstrates that the polysulfide species can be efficiently detected and the major

polysulfide species observed in (-) ESI/MS spectra is the radical species S3-. Similar observation is

also reported by other groups [21]. However, analysis of the polysulfide species in the electrolyte of

Li-S batteries by direct injection (-) ESI/MS is hindered by the high concentration of the supporting

salt LiTFSi. The major limitation for ESI/MS is that the ionization process is susceptible to ion

suppression, which mainly results from the competition between matrix components and the

analytes. In a Li-S electrolyte, the concentration of LiTFSi is much higher than the concentration of

polysulfides. The high concentration of LiTFSi competes with the polysulfide species during the

ESI process; therefore the ionization of polysulfides will be suppressed. Figure 1B shows HPLC

(-)ESI/MS spectra of the electrolyte recovered from the Li-S battery. It clearly indicates that the

recorded ions during ESI/MS scan were mainly the TFSi anion with m/z ratio 280, while

polysulfide ions were greatly suppressed by the TFSi anion and became undetectable. In order to

detect polysulfides in the electrolyte of a Li-S battery with ESI/MS, they have to be separated from

the high concentration of TFSi anions. However, ionic polysulfide species cannot be simply

separated by RP-HPLC system due to the lack of retention mechanism [18].


4
The derivatization reaction of polysulfide ions is shown in equation 2. By bonding with two

4-(dimethylamino) benzoyl groups, the polysulfide ions were derivatized into covalent compounds

which have retention in a RP-HPLC column. To ensure the complete derivatization of polysulfide

species and minimize the potential disproportion reaction by dilution of polysulfide solution, excess

amount of derivatization reagent was used. Typically, 250ul polysulfide solution was mixed with

500ul 65.2mM derivatization reagent solution. Figure 2 clearly demonstrates that the effective

separation of polysulfide species derivatized by 4-(dimethylamino) benzoyl chloride was

successfully achieved. A logarithmic dependence of retention time on the chain length of

methylated polysulfide species was reported in ref 21. The same relationship was found for the

polysulfide species derivatized by 4-(dimethylamino) benzoyl chloride, as shown in the inset in

figure 2.

As shown in equation 2, the derivatization of polysulfide species introduces the 4-(dimethylamino)

benzoyl functional group. The basic tertiary amine is a good proton acceptor and thus the

derivatized polysulfide species can be efficiently ionized during ESI process. Figure 3 shows the

(+)ESI/MS spectra of [RS3R+H]+ as an example, in which not only the expected m/z ratio but also

the unique isotopic distribution from 32S and 34S were observed.

ESI/MS-MS scans on the observed parent ions were performed for further assurance. Figure

4 shows the collision-induced dissociation (CID) spectra of parent ions with m/z 393 ([RS3R+H]+).

During the CID process, the carbon-sulfur bond in the derivatized polysulfide species was broken

forming daughter ions of 4-(dimethylamino) benzoyl cation with m/z 148. The mechanism is shown

in the inset in figure 4. Based on the ESI/MS and ESI/MS-MS data, it can be confidently concluded

5
that the polysulfide species can be efficiently derivatized by 4-(dimethylamino) benzoyl chloride

and the derivatized polysulfide species can also be effectively separated and detected by RP-HPLC

(+)ESI/MS.

Compared to the derivatization methods reported in the literature [21, 22, 24], the method

reported in this work has two distinct advantages: unique identification and good separation. For

example, Barchasz et al [24] used HPLC/UV to determine the dissolved polysulfides during the

discharge of a Li-S battery by derivatizing the dissolved polysulfide species with methyl

trifluoromethanesulfonate. The polysulfide species were hard to be qualitatively identified due to

the similarity of the UV-Vis spectra for each methylated polysulfide species. The identification of

each derivatized species was solely based on the retention time of the chromatographic peaks. One

of the potential uncertainties for the analysis was the co-elution. Indeed, their results [23] show the

co-elution of derivatized (CH3)2S8 and elemental sulfur at around 70 min. In the current work, the

identification of derivatized polysulfide species was based on retention time, ESI/MS, and ESI/MS-

MS, which almost excluded any potential mis-identification. In addition, as shown in figure S-2

(supporting information), the baseline separation of the derivatized species and elemental sulfur (S8)

was achieved with the novel derivatization method.

Figure 5 shows The RP-HPLC ESI/MS analysis for the electrolyte recovered from the

discharged Li-S cell. The derivatized polysulfides in the electrolyte with high concentration LiTFSi

were clearly separated in the RP HPLC column and effectively detected by ESI/MS. It’s also

noticeable in figure 5 that there are some doublet peaks in chromatograms for the derivatized

polysulfide species with long sulfur chain length. This phenomenon is due to the so-called

“injection solvent effect”. When an analyte is more soluble in the solvent in the injected sample than

the mobile phase, the chromatographic peak could be distorted resulting peak splitting or

broadening. Since DME in the injected samples is a more non-polar solvent than the mobile phase,

the chromatographic peak splitting is seen in figure 5. Such peak splitting can be eliminated or

minimized under gradient HPLC condition. As a comparison, peak splitting is not evident in figure
6
2, because the polarity of acetonitrile in injected samples was similar to that of the mobile phase.

Table 1 tabulates the comparison of polysulfides distribution in solution A, B and the electrolyte

recovered from the discharged Li-S cell.

A couple of interesting observations are worth noting. First, the polysulfide specie with

certain sulfur chain lengths cannot be synthesized by simply mixing stoichiometric elemental sulfur

and S2-. As shown in figures 1 and 2, instead of S42-, almost a full spectrum of polysulfides were

detected in the solution made by mixing 3:1 mole ratio of elemental sulfur and Na2S and the

distribution of the polysulfides was solvent dependent, as shown in table 1. Secondly, the major

polysulfide species in the solutions A and B, and electrolyte recovered from a discharged Li-S cell

were the polysulfide species with three sulfurs. It may indicate that the polysulfide species with

three sulfur in either di-anion form (S32-) or radical form (S3-) in the discharged electrolyte have

high solubilty and stablity in electrolyte , especially the radical species S3- which has been found in

recent study using electron paramagnetic resonance [23].

By using a novel derivatization reagent with polar and basic functional groups, the

polysulfide species in an organic electrolyte containing high concentration of salt can be

successfully separated and analyzed by the RP-HPLC ESI/MS. For the first time, the relative

distribution of polysulfide species in such electrolyte was quantitatively and reliably determined.

The HPLC ESI/MS assay will provide a reliable tool for the analysis of polysulfide species in the

Li-S battery, which is critical for the investigation of the sulfur redox mechanism.

Experimental Section

Elemental sulfur (Fisher Scientific ), lithium metal, sodium sulfide, ammonium formate, anhydrous

acetonitrile, HPLC grade methanol, HPLC grade water, 4-(dimethylamino) benzoyl chloride (all

from Sigma Aldrich), and dimethoxyethane, lithium bis(trifluoromethane) sulfonimide (FERRO)

were purchased with the most adequate grades and used without further treatment.

Two polysulfide solutions were made by dissolving sodium sulfide (Na2S, 0.2mmol) and

elemental sulfur (0.075mmol) in dimethoxyethane (DME, 5ml) or acetonitrile (AN) in a glove box
7
filled with Ar. They are labelled as solution A and B, respectively. Both solutions were in dark

brown color.

A 3-electrode electrochemical cell was assembled. The cathode was made by a sulfur loaded

porous carbon (60 wt% sulfur); metallic Li was used as both reference and counter electrode. 1M

lithium bis(trifluoromethane) sulfonimide (LiTFSi) DME solution was used as electrolyte. The cell

was then discharged at constant current of 0.2 mA from 3.37 V (OCP vs. Li) to 1.75 V. 10 mL

electrolyte was then recovered from the discharged cell.

The derivatization agent was made by dissolving 4-(dimethylamino) benzoyl chloride

(roughly 6 mg) in pure solvent (either DME or AN, 500 µL ) in a 1.5 mL sample vial; the

polysulfide samples (250 µl ) were then added and thoroughly mixed with the derivatization agent

with matching solvent for two hours. The entire derivatization processes were done in the Ar-filled

glove box. The samples were then sealed with PTFE septum and taken out for HPLC-ESI/MS

analysis. The details for HPLC-MS instrument and experiment conditions are described in the

supplemental material.

Acknowledgements
The authors from UMB and BNL are indebted to the Assistant Secretary for Energy Efficiency and
Renewable Energy, Office of Vehicle Technologies of the U.S. Department of Energy for financial
support under Contract Number DEAC02-98CH10886. The author from WUT is grateful for the
partially supported by the Fundamental Research Funds for the Central Universities China.

Received: ((will be filled in by the editorial staff))


Revised: ((will be filled in by the editorial staff))
Published online: ((will be filled in by the editorial staff))

8
[1] P.G. Bruce, S.A. Freunberger, L.J. Hardwick, J-M Tarascon, Nat. Mater. 2012, 11, 19.

[2] L.F. Xiao, Y.L. Cao, J. Xiao, B. Schwenzer, M.H. Englhard, L.V. Saraf, Z. Nie, G.J.

Exarhos, J. Liu, J. Mater. Chem. A 2013, 1, 9517.

[3] X.L. Ji, L.F. Nazar, J. Mater. Chem. 2010, 20, 9821.

[4] W. Giggenbach, J. Inorg. Nucl.Chem.1968, 30, 3189.

[5] R.P. Marin, W.H. Doub, J.L. Roberts, D.T. Sawyer, Inorg. Chem. 1973, 12, 1921.

[6] R.D. Rauh, F.S. Shuker, J.M. Marston, S.B. Brummer, J. Inorg. Nucl. Chem. 1977, 39, 1761.

[7] R. Bonnaterre, G. Cauquis, J. Chem.Soc., Chem.Commun. 1972, 1, 293.

[8] F. Seel, H.J. Guttler, G. Simon, A. Wieckowski, Pure and Appl. Chem. 1977, 49, 45.

[9] M.U.M. Patel, R. Demir-Cakan, M. Morcrette, J-M. Tarascon, M. Gaberscek, R. Dominko,

ChemSusChem. 2013, 6, 1177.

[10] G.L. Guillanton, Q.T. Do, D. Elothmaini, J. Electrochem. Soc. 1996, 143, L223.

[11] J. Badoz-Lambling, R. Bonnaterre, Electrochim. Acta 1978, 21, 119.

[12] B.S. Kim, S.M. Park, J. Electrochem. Soc. 1993, 140. 115.

[13] J. Paris, V. Plichon, Electrochim. Acta 1981, 26, 1823.

[14] C. Barchasz, F. Molton, C. Duboc, J-C Lepretre, S. Patoux, F. Alloin, Anal. Chem. 2012, 84,

3973.

[15] M. Hagen, P. Schiffels, M. Hammer, S. Dorfler, J. Tubke, M.J. Hoffmann, H. Althues, S.

Kaskel, J. Electrochem. Soc. 2013, 160, A1205.

[16] M.A. Lowe, J. Gao, H.D. Abruna, RSC Advances 2013, 3, 1957.

[17] Y. Diao, K. Xie, S.H. Xiong, X.B. Hong, J. Electrochem. Soc. 2012, 159, A421.

[18] D. Zheng, D.Y. Qu, J. Electrochem. Soc. 2014, 161, A1164

[19] Z. Uddin, R. Markuszewski, D.C. Johnson, Anal. Chim. Acta 1987, 200, 115.

[20] R. Steudel, G. Holdt, T. Göbel, J. Chromatogr. A 1989, 475, 442.


9
[21] A. Kamyshny, A. Goifman, J. Gun, D. Rizkov, O. Lev, Environ. Sci. Technol. 2004, 38,

6633.

[22] A. Kamyshny, I. Ekeltchik, J. Gun, O. Lev, Anal. Chem. 2006, 78, 2631.

[23] M. Vijayakumar, N. Govind, E. Walter, S. D. Burton, A. Shukla, A. Deveraj, J. Xiao, J. Liu,

C. M. Wang, A. Karim, S. Thevuthasan, Phys.Chem. Chem.Phys. 2014, 16, 10923.

[24] C. Barchasz, F. Molton, C. Duboc, J. C. Leprêtre, S. Patoux, F. Alloin, Anal. Chem. 2012,

84, 3973.

10
Figure 1. (A) (-) ESI/MS spectra of solution B ( Na2S and S8in DME, molar ratio of Na2S:S is 1:3)
through direct infusion. (B) (-)ESI/MS spectra of electrolyte sample from a discharged Li-S battery
by HPLC (-)ESI/MS.

Figure2. The chromatograms of derivatized solution A (Na2S and S8in acetonitrile, molar ratio of
Na2S:S is 1:3). A=total ion current chromatogram; B= ion chromatogram for m/z 329; C= ion
chromatogram for m/z 361; D= ion chromatogram for m/z 393; E= ion chromatogram for m/z 425;
F= ion chromatogram for m/z 457; G= ion chromatogram for m/z 489; H= ion chromatogram for
m/z 521; I= ion chromatogram for m/z 553. Inset: The dependencies of the retention time on
polysulfide chain length for the derivatized RSnR.
11
Figure 3. The ESI/MS spectrum of protonated derivatized polysulfide [RS3R+H]+.

Figure 4. The CID spectrum of protonated derivatized polysulfide [RS3R+H]+.

12
Figure5. The chromatograms of derivatized recovered electrolyte from discharged Li-S battery.
A=total ion current chromatogram; B= ion chromatogram for m/z 329(not observed); C= ion
chromatogram for m/z 361; D= ion chromatogram for m/z 393; E= ion chromatogram for m/z 425;
F= ion chromatogram for m/z 457; G= ion chromatogram for m/z 489; H= ion chromatogram for
m/z 521; I= ion chromatogram for m/z 553. Inset: The ESI/MS spectra of protonated derivatized
polysulfide [RS3R+H]+.

Table 1. HPLC ESI/MS results of solution A, B and the electrolyte recovered from the discharged
Li-S cell. The relative intensity is obtained by normalizing the peak area of each ion chromatogram
to the ion chromatogram with the highest peak area for each electrolyte. Each electrolyte was
analyzed in triplicate. The % value in brackets is the relative standard deviation for the triplicate
measurements.

[M+H]+ Retention Relative Intensity Relative Intensity Relative Intensity


M Time for solution A for solution B for the electrolyte
R2 S 329 6.19±0.12 1.000(0.4%) 1.000(15.2%) 0
R2S2 361 7.02±0.19 0.253(6.0%) 0.129(16.2%) 0.352(4.8%)
R2S3 393 8.13±0.02 0.854(1.1%) 0.539(10.2%) 1.000(6.4%)
R2S4 425 10.09±0.05 0.195(2.8%) 0.142(17.7%) 0.229(5.6%)
R2S5 457 11.95±0.02 0.194(4.0%) 0.177(17.0%) 0.278(7.0%)
R2S6 489 15.96±0.06 0.116(5.3%) 0.120(19.3%) 0.181((8.1%)
R2S7 521 20.66±0.07 0.035(4.1%) 0.052(23.4%) 0.121(9.7%)
R2S8 553 28.57±0.17 0.002(13.1%) 0.017(15.8%) 0.081(8.1%)

13
The polysulfide species dissolved in aprotic solutions can be separated and analyzed by RP-HPLC

ESI/MS. For the first time, the relative distribution of polysulfide species in the electrolyte

recovered from Li-S batteries was quantitatively and reliably determined.

Keyword: polysulfide, Li-S battery, HPLC ESI/MS

Dong Zheng, Deyu Qu, Xiao-Qing Yang,and Deyang Qu*


Title: Quantitative and Qualitative Determination of Polysulfide Species in the Electrolyte of
Lithium-Sulfur Battery by HPLC ESI/MS with One-step Derivatization

ToC figure :

14
Copyright WILEY-VCH Verlag GmbH & Co. KGaA, 69469 Weinheim, Germany, 2013.

Supporting Information

Title: Quantitative and Qualitative Determination of Polysulfide Species in the Electrolyte of


Lithium-Sulfur Battery by HPLC ESI/MS with One-step Derivatization

Dong Zheng, Deyu Qu, Xiao-Qing Yang,and Deyang Qu*

1, Details for HPLC-MS

An Agilent 1100 isocratic pump (from Agilent) with a 7725i Rheodyne manual injector was used to

deliver methanol/water mixture (90:10,v/v, with 5mM ammonium formate) through a Luna HPLC

column (from Phenomenex, C18, 4.6*250mm, 5um) at flow rate 0.75mL/min. The injection volume

is 20 µL. A split T was used to introduce about 10% total flow from HPLC into the mass

spectrometer.

A Quattro LC mass spectrometer (triple quadrupole, from Micromass-Waters) with ESI

source was operated under positive mode. For (+) ESI/MS, the typical source parameters were set as

following: Capillary 2.50kV, Cone 20V, Extractor 5 V, RF lens 0.1V, Source temperature 100°C,

Desolvation temperature 250°C, nitrogen gas flow is 90L/hr for Nebuliser and 600L/hr for

Desolvation. For HPLC (+)ESI/MS run, the m/z ratio from 200 to 1000 was recorded; for HPLC

(+)ESI/MS-MS (daughter ion scan) run, the m/z ratio from 20 to 1000 was recorded, and collision

voltage was set at 20, and argon as collision gas. For (-) ESI/MS, the typical source parameters were

set as following: Capillary 3.50kV, Cone 30V, Extractor 5 V, RF lens 0.1V, Source temperature

90°C, Desolvation temperature 200°C, nitrogen gas flow is 50L/hr for Nebuliser and 500L/hr for

Desolvation. For both HPLC (-)ESI/MS and (-)ESI/MS by direct infusion , the m/z ratio from 20 to

600 was recorded. A syringe pump was used to introduce the simulated electrolyte into MS at a

flow rate around 50ul/min.

15
All of the electrochemical measurements were carried out using an Arbin MSTAT (Arbin

instruments) controlled by MITS PRO software.

2, HPLC-MS results

Figure S-1. Comparison of Chromatograms for HPLC/UV (A), HPLC/ESI-MS (B, positive mode;

C, negative mode) for the methylated polysulfide species from our preliminary results. No

chromatographic peaks are observed in HPLC/ESI-MS after 4 minute where the methylated

polysulfide species are eluted out in HPLC-UV.

16
Elemental Sulfur,
24.84min
1.4

RS7R, RS8R,
1.2 20.46min 28.74min

1.0
Relative Intensity

0.8

0.6

0.4

0.2

0.0
15 20 25 30 35
Retention time (min)

Figure S-2. Comparison of Chromatographic separation for derivatized polysulfide species RS7R (in

red), RS8R (in blue) and elemental sulfur (in black). Noted that the chromatogram of elemental

sulfur is from HPLC/UV and the chromatograms of derivatized polysulfide species are from

HPLC/ESI-MS.

17

You might also like