You are on page 1of 14

Supporting information

Green modification of Graphene oxide nanosheets under

specific pH conditions

R. Castellanos-Espinoza1, S. Fernández-Tavizón2, U. Sierra-Gómez2, E. A. Elizalde-Peña3, G.


Luna-Bárcenas4, L. A. Baldenegro-Pérez5, Lilian I. Olvera6, L.V. G onzález-Gutiérrez1, C. M.
Ramos-Castillo3, Noé Arjona1, B. L. España-Sánchez1*

1
Centro de Investigación y Desarrollo Tecnológico en Electroquímica CIDETEQ SC. Parque
Tecnológico Querétaro s/n, Sanfandila, Pedro Escobedo, Querétaro. C.P. 76703, México.
Laboratorio Nacional de Materiales Grafénicos, Centro de Investigación en Química Aplicada CIQA,
2

Blvd. Enrique Reyna Hermosillo, San José de los Cerritos, Apdo. 140, Saltillo, Coahuila. C.P. 25294,
México.
Dirección de Investigación y Posgrado de la Facultad de Ingeniería, Universidad Autónoma de
3

Querétaro, Santiago de Querétaro, C.P. 76140, México.


CINVESTAV del IPN Unidad Querétaro, Libramiento Norponiente No. 2000, Fracc. Real de Juriquilla,
4

Querétaro. C.P. 76230, México.


Centro de Ingeniería y Desarrollo Industrial (CIDESI), Av. Playa Pie de la Cuesta 702, Desarrollo San
5

Pablo, Querétaro. C.P. 76130, México.


6
Instituto de Investigaciones en Materiales, Universidad Nacional Autónoma de México, Apartado postal
70-360, C.U. Coyoacán, Ciudad de México C.P. 04510, México.

Corresponding author: B.L. España-Sánchez


lespana@cideteq.mx
S1. Electrochemical characterization

Electrochemical experiments were performed to determine the electrochemically active surface

area (EASA, cm2) and specific surface area (SSA, m2/g). For this purpose, an electrochemical

cell with a three electrode configuration was employed. The reference electrode was the

saturated mercury sulfate electrode (MSE, saturated in K2SO4 E=0.65V vs. NHE), and a graphite

rod was used as the counter electrode. Glassy carbon tips for rotatory-disk electrode tests were

used as working electrodes, and they polished until mirror-finishing and electrochemically

cleaned prior its use (50 cycles in 0.5 M H 2SO4). The EASA was determined by the capacitances

method [1] using 0.5 M H2SO4 as the electrolyte. The cyclic voltammetry graphs were obtained ±

50 mV from the open circuit potential at different scan rates: 20, 40, 60, 80 and 100 mV/s. The

theoretical capacitance for carbonaceous materials was 40 µF/cm2 [2,3].

S2. Computational methods


Total energy calculations were performed within the framework of a linear combination of

atomic orbitals (LCAO) as implemented in the SIESTA code [4]. A double zeta polarized (DZP)

basis set [5] was employed. The core electrons were treated with norm-conserving

pseudopotentials in their non-local form [6]. A mesh-cutoff energy of 300 Ry is applied to

sample the electronic density in real space. Conjugate gradient method was used for structural

optimization with a maximum value in the interatomic forces of 0.01 eV/Å. Dispersion

interaction were evaluated using DFT−D as implemented in SIESTA code. The Figure S9 shows

the graphene nanosheet model used for functionalization and reaction path studies, a larger

supercell of 40 x 40 x 40 was used in order to avoid interaction between adjacent images. Python

Algorithms for Searching Transition stAtes (PASTA) was implemented for reaction path
calculation [7]. The vibrational frequencies were calculated in order to check the reliability of transition

states. The vibrational frequencies were calculated only for different adsorbed species on graphene oxide

support, for this reason, the carbon, and oxygen of GO support atoms are fixed during vibrational

calculation. A VESTA software was used for the visualization of structural models [8].

S3. Raman spectroscopy measurements of GO and GO−Lys

Density defects in GO nanosheets were determined by Raman spectroscopy, using a Horiba

Xplora at 632.8 nm. Characteristic D band attributed to the disordered regions of sp 3 domains

from the vibrational symmetry A1g (1340 cm−1) and G band associated to the vibrational

symmetry E2g of sp2 domains (1588 cm−1) were obtained (Figure S1) [9,10]. ID/IG relation was

obtained to determine the defect distance (LD) and defect density (ƞD), according to the equations:

( )
−1
2 −9 4 ID
L D (nm) =1.8 ± 0.5 ×10 λ
2
L Equation 1
IG

( )
ID
22
(1.8 ±0.5) ×10
n D ( cm−2 ) = Equation 2
λ 4L IG

To determine the ƞD condition, defect density present a LD ≥ 10.


Figure S1. Raman spectroscopy measurements of GO and GO−Lys. (a) GO, (b) GO−Lys pH 2,

(c) GO−Lys pH 9.8 y (d) GO−Lys pH 14.


Table S1. Comparison of experimental values for deconvoluted XPS spectrum of C 1s.

Bond Reported Binding Experimental


energy (eV) Binding energy (eV)
[11]
GO GO−Lys pH 2 GO−Lys pH 9.8 GO−Lys pH 14
C═C 284.4±0.1 284.5 284.5 284.5 284.5
C−C 285.1±0.1 285.0 285.1 285.1 285.1
C−OH, C−N 285.9±0.1 285.9 285.9 285.9 285.9
C−O−C 286.5±0.3 286.8 286.8 286.6 286.8
C═O 287.7±0.5 287.5 287.6 287.7 288.0
C(═O)−O 288.8±0.2 288.6 288.8 288.8 288.8

Figure S2. High-resolution deconvoluted XPS spectra of N1s core levels of a) GO, b) GO−Lys

pH 2, c) GO−Lys pH 9.8, and d) GO−Lys pH 14.


Figure S3. High-resolution deconvoluted XPS spectra of O1s core levels of a) GO, b) GO−Lys

pH 2, c) GO−Lys pH 9.8, and d) GO−Lys pH 14.


S4. Surface area by the BET method

The surface area calculation was carried out using the Brunauer-Emmett-Teller (BET) method,

which was obtained by absorption/desorption of nitrogen using equipment Quantachrome S-BET

Autosorb iQ2. Before performing the analysis, the samples were dried at 150 o C for 24 h.

Table S2. Surface characterization of GO nanosheets and their modification with Lys under

specific pH conditions.

Surface area Electrochemically Improvement SSA

BET surface reactive factor (m2/g)

(m2/g) area (cm2)

GO 3.77 2.75 14.05 241.79

GO−Lys pH 2 4.36 1.91 9.77 168.27

GO−Lys pH 9.8 5.04 9.35 47.73 821.35

GO−Lys pH 14 4.55 2.23 11.37 187.52


0.020

0.015
a b
0.010

0.005

i/mA
0.000

-0.005

-0.010 GO-20 mV
GO-40 mV
-0.015 GO-60 mV
GO-80 mV
GO-100 mV
-0.020
0.10 0.12 0.14 0.16 0.18 0.20

E (V)

c d

e f

g h

Figure S4. Cyclic voltammetry measurements of a) GO, c) GO−Lys pH 2, e) GO−Lys pH 9.8, g)

and GO−Lys pH 14 in the solution of 0.5 M H2SO4, with scan rate of 20-100 mV/s. Variation of

anodic and cathodic peak currents of b) GO, d) GO−Lys pH 2, f) GO−Lys pH 9.8, h) and

GO−Lys pH 14
Figure S5. Lysine behavior under specific pH conditions.

Figure S6. Electrostatic attraction of protonated amines of Lys with GO nanosheets through the

cation-π interactions at pH 2.
Figure S7. Reaction mechanism proposed for GO with Lys at pH 2 (esterification reaction).
Figure S8. Reaction mechanism proposed for GO with Lys at pH 9.8 (C−N bond formation).
Figure S9. Graphene nanosheet model used for template for functionalization and reaction path

studies.

Figure S10. Schemes a) and c) show the two stages and the main reaction steps for the formation
of the ester bond. b) and d) calculated reaction path (NEB method) for the formation of the ester
bond.
Figure S11. Scheme a) shows the main reaction steps for the formation of the C-N bond. b)
calculated reaction path (NEB method) for the formation of the C-N bond.
References

[1] N. Arjona, M. Guerra-Balcázar, L. Álvarez-Contreras, Hierarchical Pd and Pt structures


obtained on 3D carbon electrodes as electrocatalysts for the ethylene glycol electro-
oxidation, Appl. Surf. Sci. 571 (2022) 151246.
https://doi.org/10.1016/J.APSUSC.2021.151246.
[2] A. Roy, A. Ray, S. Saha, M. Ghosh, T. Das, M. Nandi, G. Lal, S. Das, Influence of
electrochemical active surface area on the oxygen evolution reaction and energy storage
performance of MnO2-multiwalled carbon nanotube composite, Int. J. Energy Res. 45
(2021) 16908–16921. https://doi.org/10.1002/ER.6885.
[3] H. Han, J. Won Paik, M. Ham, K. Min Kim, J. Kuen Park, Y. Kyu Jeong, H.S. Han, J.W.
Paik, M.J. Ham, K.M. Kim, Y.K. Jeong, J.K. Park, Atomic Layer Deposition-Assisted
Fabrication of Co-Nanoparticle/N-Doped Carbon Nanotube Hybrids as Efficient
Electrocatalysts for the Oxygen Evolution Reaction, Small. 16 (2020) 2002427.
https://doi.org/10.1002/SMLL.202002427.
[4] J.M. Soler, E. Artacho, J.D. Gale, A. García, J. Junquera, P. Ordejón, D. Sánchez-Portal,
The SIESTA method for ab initio order-Nmaterials simulation, J. Phys. Condens. Matter.
14 (2002) 2745. https://doi.org/10.1088/0953-8984/14/11/302.
[5] D. Sánchez-Portal, J. Junquera, Ó. Paz, E. Artacho, Numerical atomic orbitals for linear-
scaling calculations, Phys. Rev. B. 64 (2001) 235111.
https://doi.org/10.1103/PhysRevB.64.235111.
[6] L. Kleinman, D.M. Bylander, Efficacious Form for Model Pseudopotentials, Phys. Rev.
Lett. 48 (1982) 1425. https://doi.org/10.1103/PhysRevLett.48.1425.
[7] S. Kundu, S. Bhattacharjee, S.C. Lee, M. Jain, PASTA: Python Algorithms for Searching
Transition stAtes, Comput. Phys. Commun. 233 (2018) 261–268.
https://doi.org/10.1016/J.CPC.2018.06.026.
[8] K. Momma, F. Izumi, VESTA 3 for three-dimensional visualization of crystal, volumetric
and morphology data, urn:issn:0021-8898. 44 (2011) 1272–1276.
https://doi.org/10.1107/S0021889811038970.
[9] A.C. Ferrari, Raman spectroscopy of graphene and graphite: Disorder, electron–phonon
coupling, doping and nonadiabatic effects, Solid State Commun. 143 (2007) 47–57.
https://doi.org/10.1016/J.SSC.2007.03.052.
[10] S. Eigler, A. Hirsch, S. Eigler, A. Hirsch, Chemistry with Graphene and Graphene Oxide
—Challenges for Synthetic Chemists, Angew. Chemie Int. Ed. 53 (2014) 7720–7738.
https://doi.org/10.1002/ANIE.201402780.
[11] B. Lesiak, L. Kövér, J. Tóth, J. Zemek, P. Jiricek, A. Kromka, N. Rangam, C sp2/sp3
hybridisations in carbon nanomaterials – XPS and (X)AES study, Appl. Surf. Sci. 452
(2018) 223–231. https://doi.org/10.1016/J.APSUSC.2018.04.269.

You might also like