You are on page 1of 29

Journal Pre-proof

Theoretical and experimental studies of tin electrodeposition

S. Bakkali , M. Cherkaoui , A. Boutouil , M.R. Laamari ,


M. Ebn Touhami , M. Belfakir , A. Zarrouk

PII: S2468-0230(19)30696-0
DOI: https://doi.org/10.1016/j.surfin.2020.100480
Reference: SURFIN 100480

To appear in: Surfaces and Interfaces

Received date: 30 November 2019


Revised date: 23 January 2020
Accepted date: 15 February 2020

Please cite this article as: S. Bakkali , M. Cherkaoui , A. Boutouil , M.R. Laamari , M. Ebn Touhami ,
M. Belfakir , A. Zarrouk , Theoretical and experimental studies of tin electrodeposition, Surfaces and
Interfaces (2020), doi: https://doi.org/10.1016/j.surfin.2020.100480

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


Theoretical and experimental studies of tin
electrodeposition
S. Bakkali1,2, M. Cherkaoui2, A. Boutouil1, M.R. Laamari1, M. Ebn Touhami3, M.
Belfakir3, A. Zarrouk4,*
1
Laboratoire de Chimie Analytique et Moléculaire (LCAM), Université Cadi Ayyad, Faculté
Polydisciplinaire, Sidi Bouzid, B.P. 4162, 46 000 Safi, Morocco.
2
Laboratoire de Matériaux d’Electrochimie et d’Environnement, Faculté des Sciences, Université Ibn
Tofaïl,14000Kénitra, Morocco.
3
Laboratory of Materials Engineering and Environment: Modeling and Application, Faculty of
Sciences, University Ibn Tofail, PO Box. 133 – 14000, Kénitra, Morocco.
4
Laboratory of Materials, Nanotechnology and Environment, Mohammed V University, Faculty of
Sciences, P.O. Box. 1014, Rabat, Morocco.
────────────────
* Corresponding author. Tel.: +212 665 201 397. Fax.: +212 537 774 261.
E-mail address: azarrouk@gmail.com (Abdelkader ZARROUK).
Abstract

Tin electrodeposition from acid solutions consumes less electrical energy and produces bright

Sn plating. However, the coating obtained has poor quality. Consequently, organic additives

are added to improve the effeteness of deposit. In the current article, the effect of decyl

glucoside, as a green additive, on tin electrodeposition in the acid sulphate medium was

studied by the combination of theoretical and experimental approaches. The polarization

curves indicate that the tested additive is not electroactive. Likewise, this compound modifies

the current density and the activation energy of tin reduction as well as the hydrogen

evolution and the apparent diffusion coefficient. The additive does not affect the electron

transfer during the electrodeposition process. The fraction of the electrode surface covered by

the adsorbed additive molecules, at the peak potential, is 0.545. The reduction of Sn2+ ions is

diffusion-controlled reaction. SEM and XRD analyses show that the additive improves the

quality and change the preferred orientation of the tin deposited. The most relevant quantum

chemical parameters of the additive were also calculated. The Parr functions and molecular

electrostatic potential plot show that the heterocyclic ring is the most active part of the

molecule. In addition, molecular dynamics simulation reveals the optimized adsorption

configuration of the tested molecule on the iron surface and emphasizes the spontaneity and

1
the physical nature of the additive adsorption. Theoretical studies are in good agreement with

the experimental results.

Keywords: Tin; Electrodeposition; Linear sweep voltammetry; SEM; XRD; Quantum

chemical calculations.

1. Introduction

Tin and tin alloy coatings are of great interest in electrometallurgy. In fact, tin is environment

friendly and the tin coatings are corrosion resistant, non-toxic and ductile. Similarly, they

have good electrical conductivity and high electrical reliability[1]. Tin and tin alloys plates

are often used in electronics, food equipment, lithium-ion battery and in the field of renewable

energy [2-6]. Recently, Tin-Based Electrocatalysts, used in the reduction of CO2 have gained

growing interest [7 8]. Tin plat is usually carried out from acidic or alkaline solutions. The

alkaline electrolytes contain tin(IV) in the form of hydroxystannate or complexes with di- and

tri-phosphate ions[9]. However, alkaline baths have certain difficulties; they require high

temperatures and must not contain Sn2+ ions. In the acidic bath, divalent tin is the main metal

cation; this type of bath contains sulphate, fluoroborate, or methane sulfonate[10]. From

acidic baths, tin is strongly deposited. Therefore, the deposits obtained are not satisfactory.

Thus, different organic compounds were added to the tinning bath to improve the quality of

the deposits [11-16]; this has made the solutions complicate and difficult to control. Note that,

the control of the tinning bath enlarges the perspective of the tin electrodeposition application

in many fields. The aim of this study lies in the use of a simple and easy to control plating

solution working at the ambient temperature. On the other hand, give an insight into the

mechanism of the tin electrodeposition in the absence and presence of the decyl glucoside

(Fig. 1) by the combination of theoretical and experimental studies. The tested additive is a

plant based non-ionic surfactant. It is highly safe compound, used in skin care products. The

experimental study was performed by voltammetry and the characterization of the deposit was

2
made by Scanning Electron Microscopy and XRD. The theoretical approach was used in the

framework of the density-functional theory and Molecular dynamic simulations.

2. Experimental Procedure

2.1. Electrochemical measurements

The electrolysis cell was a borosilicate glass (Pyrex®) cylinder closed by a cap with five

apertures. Three of them were used for the electrodes. Linear sweep voltammetry

measurements were performed using an iron disk (1 cm2) as a working electrode, Pt wire as a

counter and a saturated calomel electrode (SCE) as a reference electrode. Deposits were

carried out with a tin plate (1.34 cm2) counter electrode. The experimental procedure of this

part was identical to previously published work [17]. The compositions of the plating baths

are listed in Table 1. The temperature has been set at 298 ± 2 K.

2.2. Computational Details

Quantum chemistry calculations were performed using the Gaussian 09 program package [18]

at 298 K with water as a solvent in the CPCM theoretical model. All calculations were carried

out using the B3LYP method and the 6-311G++ (d,p) basis set. In order to better understand

the electronic interaction between the studied additive molecule and iron surface several

theoretical parameters were evaluated in the aqueous phase. The evaluated parameters include

the EHOMO, ELUMO gap energy between ELUMO and EHOMO (∆E), global hardness (η),

electronegativity (χ) and dipole moment (μ) according to [19].

E   ELUMO  EHOMO  ( 1)

1
  ELUMO  EHOMO  (2)
2

1
   ELUMO  EHOMO  (3)
2

The parr functions are obtained by analysing the Mulliken atomic spin density (ASD) at the

radical anion  sra  r  and at the radical cation  src  r  . These functions play a vital role in the

3
identification of the most favourable electrophilic ( Pk ) and nucleophilic ( Pk ) sites. The Parr

functions are given by the following equations [20]:

Pk  r    src  r  For electrophilic attacks (4)

Pk  r    sra  r  For nucleophilic attacks (5)

Where  src  r  and  sra  r  represent the atomic spin density of the radical cation and the

radical anion, respectively.

The local electrophilicity ωk and the local nucleophilicity Nk were obtained according to:

k   Pk (6)

Nk  NPk (7)

Where, ω and N are the electrophilicity and nucleophilicity indices, respectively.

The molecular dynamics simulations were carried out in a simulation box with periodic

boundary conditions using Materials Studio (available from BIOVIA Company). Fe (110)

was chosen as the simulated surface as it is considered as the most stable plane [21]. The final

 3
simulation cell (29.78 × 29.78 × 47.18 A ) was created with a system, which contains 500

molecules of water and a single additive molecule. The temperature of the system was kept

constant using Andersen’s method. All the simulations were performed under the NVT

canonical ensemble. The simulation time was 500 ps with a time step of 1 fs. The trajectory

information was recorded once every 1000 steps. The COMPASS force field has been utilized

to optimize all the structures of the system considered. The adsorption energy and the binding

energy were calculated under equilibrium conditions, according to Eqs:

Eadsorption  Etotal   Esurface  Eadditive  (8)

Ebinding   Eadsorption (9)

Where Etotal is the total energy of the entire system; Esurface referred to the total energy of Fe

(110) surface and Eadditive represents the total energy of the additive.

4
3. Results and discussion

3.1. Linear sweep voltammogram

The voltammogram, recorded in the solution containing only the additive, shows that this

additive has no electrochemical activity in the explored potential range (Fig. 2a). In the

sulfuric acid solution, the hydrogen reduction starts from about -0.67 V/SCE (Fig. 2b). When

the examined compound is added to this acidic solution, the hydrogen evolution is

considerably reduced and shifted to a more negative potential (Fig. 2c) which indicates that

the tested additive retards the hydrogen reduction [22]. Fig. 3a shows the linear sweep

voltammogram of tin electrodeposited onto the iron surface from the additive-free solution at

10 mV/S. The potential applied to the electrode was ramped down from the open circuit

potential (OCP) to -0.9 V/SCE. According to this figure, the reduction of Sn2+ ions to Sn

metal starts approximately at -0.44 V/SCE, then a current peak appears at -0.555 V/SCE. This

behaviour is attributed to the drop in Sn2+ concentration at the electrode surface. Next the

current becomes constant over a relatively large potential region, indicating that the cathodic

process is controlled by diffusion. At more negative potentials, the current is higher due to the

hydrogen reduction [23]. In presence of 10-3 mol dm-3 of the studied additive, a positive effect

on the stability of the plating solution is observed. We note also that the reduction peak

potential remains at -0.555 V/SC(Fig. 3b)E, showing that the additive does not change

the tin deposition potential [10,24]. This outcome suggests that no complex is formed with

stannous ion [25-27]. The existence of a single peak in both electrolytes indicates that the

reduction of tin ions, in the absence and presence of the additive, occurs in one elementary

two-electron step (eq10) [26-28]:

Sn 2+ + 2e  Sn (10)

Likewise, the additive decreases the reduction current [15,29,30] and negatively shifts the

potential of the hydrogen evolution. These effects could be related to the adsorption of the

additive on the metal surface [31-36], which decreases the active surface area of the working

5
electrode. To evaluate the fraction of the electrode surface area (θ) covered by the additive,

the equation (11) [22,24,37] was used:

i0  iad
 (11)
i0

Where, i0 and iad are current densities at a given potential respectively in the absence and

presence of the additive and θ is the surface coverage. The surface coverage at the peak

potential was approximately 0.545.

3.2. Effect of the scanning rate

Figs. 4&5 respectively show the voltammograms obtained, in baths with and without the

additive, at different scan rates (10-250 mV S-1). The results show that in both solutions, there

are two points to be noted when the scan rate  increases. Firstly, the peak potential shifts

toward more negative values, which indicates that the electrochemical process is irreversible.

Secondly, the cathodic peak current density increases.

For such a system, the cathodic peak potential is given by the equation (12) [38,39].

RT    nc FD   RT
E pc  E 0'   0.78  0.5Ln  2   Ln   (12)
 nc F   k RT   2 nc F
The different parameters have the same meaning as those in the literature [38]. D is the

diffusion coefficient for Sn2+.

In both electrolytes the curves Epc = f (Ln ( )) are straight lines (Fig. 6) with slopes equal to

RT
 (eq12). From these slopes c values can be determined (Table 2). The table 2 shows
2 nc F

that the c is the same in both electrolytes. This result indicates that the electron transfer

during the tin electrodeposition process is not affected by the presence of the additive [40].

Note that other additives can affect c [36,41].

To understand the kind of control that governs the tin electrodeposition process in both

electrolytes, the cathodic peak current density (IP) versus the square root of the scan rate (ν 1/2)

are plotted in Figure 7. The plots display excellent linear relationships, indicating that the

6
mass transfer controls the reduction process in both electrolytes. The reduction peak current

density Ip for an irreversible system is given by the following equation [38,39]:

I p  2.99 105 n  nc  2 D 2 2 C
1 1 1
(13)

The different parameters of equation 13 have the same meaning so that it was reported [38].

Table 3 shows the apparent diffusion coefficients of the electroactive species in each solution.

These coefficients are evaluated from the slopes of the plots Ip versus ν1/2. The value of the

diffusion coefficient calculated from the basic bath is similar to the one mentioned in the data

[42,43].

In the presence of the additive, the apparent diffusion coefficient of stannous ions decreases

(D = 1.18×10-6 cm2/S).

3.3. Effects of the temperature

The effects of temperature on the adsorption of the additive and on the electroreduction of tin

ions were elucidated. For this reason, potentiodynamic curves I = f (E) from both electrolytes

were carried out at different bath temperatures (298-328 K) using an isothermal cell.

3.3.1 Effect of the temperature on additive adsorption

Figure 8 shows the variation of the surface coverage (θ) at the peak potential versus the

temperature. It is evident that θ decreases with the increase of the temperature, a trend that

supports the mechanism of physical adsorption of the additive on the metal surface [44,45].

The temperature effect on the surface coverage (θ) may be attributed to the desorption of

some adsorbed organic molecules from the electrode surface when the temperature increases.

In fact, the desorption predominates over the adsorption at high temperature [46,47].

Therefore, as the temperature increases, the number of the adsorbed molecules of additive

decreases and the number of active centers available for Sn2+ reduction rises.

3.3.2. Effect of the temperature on tin electrodeposition

It is known that the activation energy Ea provides information on the rate-controlling step. In

fact, Ea less than 28 kJ mol-1 indicates that the rate-controlling step is the diffusion of the

7
electroactive specie. While, values of Ea higher than 43 kJ mol-1 are related to the adsorption

of the electroactive species on the electrode surface [48,49]. The activation energy of tin

electrodeposition in the absence and presence of the additive was obtained by applying the

Arrhenius equation (14) (Fig. 9).

Ea
log I L   A (14)
2.303RT

Where Ea is the apparent activation energy, R is the universal gas constant and A is the

Arrhenius pre-exponential factor.

The results are given in Table 4. In both solutions, Ea are less than 28 kJ mol-1 which indicates

that the electrodeposition process in both electrolytes is controlled by the diffusion of the

electroactive species. The results also show that the additive increases the energy barrier of

tin reduction. In general, the increase in Ea in the presence of an additive suggests its physical

adsorption [50-52].

3.4. Characterization of the deposit

To find out the effect of the additive on the crystal structure of the film electrodeposited, SEM

and XRD characterization were carried out on samples with thickness of about 10µm of

deposit performed on iron substrate, from solutions with and without the additive.

3.4.1. Morphology of the deposits

Figure 10 shows the morphology of Sn deposits obtained by SEM from both electrolytes. In

additive free solution, the tin deposit is porous, dendritic with an irregular form and a larger

crystal size (Fig. 10a). In the presence of the additive, the deposit is bright, densely packed

with a uniform arrangement and a low grain size (Fig. 10b). The adsorption of the additive on

the electrode surface appears to retard grain growth processes.

3.5.2. Crystallographic structure

The collected X-ray diffraction (XRD) results (Fig. 11) were compared with the pattern card

No.00-004-673 given in JCPDS data base. The obtained patterns indicate that all the deposits

are a single-phase of pure Sn crystallizing in the tetragonal lattice with space group P63/m.
8
The texture coefficient TC [34]. was used to describe the effect of the additive on the

preferred orientation.

I hkl
0
TC 
I hkl
(15)
 1n   II hkl
0
hkl

Where Ihkl is the relative peak intensity of tin electrodeposits and n is the number of

reflections. The index 0 refers to the relative intensities for the standard tin powder sample

(JCPDS card No. 00-004-673).

A preferred orientation is characterized by a TC greater than 1.0 [53]. From the Fig. 12a the

film deposited from additive free solution has a strong (2 0 0) preferred orientation. In the

presence of the additive, the deposit shows a (1 1 2) preferred orientation as shown in the Fig.

12b. The particle size values of the deposits are evaluated using the Scherrer formula from the

predominant peak [54,55]. We find that the main size of the crystal decreased from

92 nm to 52 nm in the presence of the additive.

3.5. Theoretical approach

Recently, computational chemistry, as a powerful tool, has been introduced to get deeper

insights into the field of electrodeposition [56-67].

3.6.1. Quantum chemical calculations

Density functional theory (DFT) is used in chemistry to clarify the geometry, the electronic

structures, the reactivity and the dynamics of atoms and molecules. Herein, we expect that

quantum chemical calculations will allow for a more precise estimation of electronic

properties. For this purpose, various quantum parameters derived from the optimized

equilibrium structure (Fig 1) are calculated and summarized in Table 5. The HOMO and

LUMO values are often associated with the electron donating and accepting ability,

respectively. The table 5 shows that the studied compound has a relatively high value of

EHOMO (-7.42 eV) which points out its higher adsorption capacity on the iron surface. Hence,

this additive has a good inhibiting performance on the deposition process.

9
Furthermore, as shown in Fig. 13a, the HOMO region is distributed on the oxygen and carbon

atoms on the heterocyclic ring. This region corresponds to the electron-donating zone that

gives electrons to the vacant d-orbital of the metal. Thus, C and O atoms on the ring are

probably the main adsorption sites of the molecule. On the other side the LUMO (Fig. 13b) is

mainly localized on the O17 and H18 and slightly extended to the neighboring C1, C2, C3,

C19, 013 and O15 atoms. The molecular gap (∆E) is related to the reactivity of the organic

molecule towards the adsorption on the electrode surface. In fact, the literature establishes that

soft molecules are characterized by a small energy gap and have good adsorption capacity on

the metallic surfaces [68,69]. ∆E of the decyl glucoside was 7.12 eV. Other electrodeposition

additives have similar values [70,71]. The high dipole moment value μ (5.4935 Debye) points

out the strong adsorption capacity of the additive [72].

Another parameter which provides information on the reactivity of organic molecules is the

hardness η. In fact, soft molecules are more reactive. Table 5 reveals that the additive

molecule can be considered a soft molecule with a high reactivity [70]. The electrophilicity

index, ω, gives information on the nucleophilic or electrophilic nature of a molecule; it shows

the ability of the molecules to accept electrons. The high value of ω reveals the high capacity

of the additive molecule to accept electrons to form a back-donating bond [58]. A

comparative approach has been carried out using molecular electrostatic potential (MEP) (Fig.

14). This map illustrates the areas in a molecule of high and low electron density. The red

areas show the region of the most negative MEP. The blue region indicates the parts of the

most positive MEP. The green parts represent the region of the zero MEP. The obtained

results emphasize the high electron density near O11, O12, O13, O53 and C4, which are

located on the heterocyclic ring. Deeper information on sites responsible for the adsorption

behavior is provided by Parr function P(r) as local molecular reactivity. Calculated values of

the local electrophilicity ωk and nucleophilicity Nk indices and the local nucleophilic Pk and

electrophilic Pk of the studied molecule are listed in the Table 6.

10
Examination of the nucleophilic Parr functions of the studied additive molecule shows that,

O11, O15, O17 and C1, could be the most nucleophilic sites presenting the maximum values

of Pk and Nk. On the other hand, C2 is the most electrophilic site since it has the highest

values of Pk and ωk. To elucidate the adsorption process of the additive on the electrode

surface, a detailed study is carried out using molecular dynamics simulations.

2.1.2. Molecular dynamics simulations

Molecular dynamic simulations were used to investigate the interaction between organic

molecule and Fe (1 1 0) surface. The equilibrium criteria limiting the fluctuations of

temperature and energy to 5-10% [73,74] was used to assert if the model system has reached

equilibrium or not. According to Fig. 15, in the last 85ps the temperature fluctuated between

285 K and 312 K, and the fluctuation of energy was very small (< 0.5%), indicating that

equilibrium state was reached by the system. The adsorption processes of the additive at

298 K is shown in the Fig. 16. First, the organic molecule approached the iron surface

(0-5ps). Then the adsorption of the ring part of the organic compound took place gradually on

the metal surface (38ps-46ps). Subsequently, the alkyl part of the additive adsorbed

progressively on the cathode (49ps-415ps). After that, the molecule kept almost the same

position on the surface, which means that the adsorption process of the molecule was

completed (415ps- 500ps).

The equilibrium structure of the additive is parallel and close to the iron surface (Fig. 16,

500ps and 500ps top view). The negative interaction energy (-357.4 kcal/mol) emphasizes the

spontaneous character of the organic compound adsorption. In addition, the large binding

energy (357.4 kcal/mol) denotes that a strong interaction takes place between the additive

molecule and the iron surface [75].

To gain further insights into the nature of metal - additive molecule interactions, Radial

distribution function (RDF) was used. (RDFs) was calculated from the simulation trajectories

[76]. In general, a small bond length ranging in between (1Å-3.5Å) characterizes the chemical
11
bond, while the typical length for the physical bond is longer than 3.5 Å [77]. The position of

the first prominent peak in the RDF curve is the most probable bond length [78]. Based on the

RDF results (Fig.17), the first prominent peak is located at 4,48Å which reflects the physical

nature of the interaction between the additive and the electrode surface.

4. Conclusion

Electrodeposition of tin from the acidic bath in the absence and presence of decyl glucoside

was investigated. Linear Sweep voltammetry measurements indicates that the additive is not

electroactive. The tin reduction reaction, in both electrolytes, is diffusion controlled and takes

place through a 2-electron transfer pathway. The variation of the scan rate shows that the

electrochemical process is irreversible. The additive adsorption on the working electrode

surface, has profound impacts. In fact, it decreases both the reduction current and the

diffusion coefficient of Sn2+, increases the activation energy of the electrode reactions and

retards the hydrogen evolution. On the other hand, the transfer coefficient seems to be not

impacted by the additive addition. The additive adsorption process is physical in nature and

depends on the temperature. X-ray diffraction patterns confirms that the deposits from both

solutions are single-phase of pure Sn crystallizing in the tetragonal lattice. The additive

significantly improves the quality of the deposit and decreases the grain size of the film.

Furthermore, the decyl glycoside changes the preferred orientation of the coating.

Additionally, the theoretical study exhibited that the additive has a small energy gap, negative

adsorption energy and high binding energy revealing its good adsorption property. The

reactivity of the additive is mainly spread over the ring part of the molecule. The findings

obtained from the theoretical studies agree with the results of the experimental tests.

Dear Editor-in-Chief,

12
This statement is to certify that all Authors of the article “Theoretical and experimental

studies of tin electrodeposition” have been seen and approved the manuscript being

submitted. We warrant that the article is the Auhor’s original work. We warrant that the

article has not received prior publication and is not under consideration for publication

elsewhere. No conflict of interest exists, or if such conflict exists, the exact nature must be

declared. On behalf of all Co-Authors, the corresponding Author shall bear full responsibility

for the submission.

All Authors agree that author list is correct in its content and order and that no modification to

the author list can be made without the formal approval of the Editor-in-Chief, and all

Authors accept that the Editor-in-Chief’s decisions over acceptance or rejection.

Corresponding author

Prof. Dr. Abdelkader ZARROUK

References
[1] M.J. Bureka, A.S. Budiman, Z. Jahed, N. Tamura, M. Kunz, S. Jin, S.M.J. Han, G. Lee, C.
Zamecnik, T.Y. Tsui, Fabrication, microstructure, and mechanical properties of tin
nanostructures, Mater. Sci. Eng., A 528 (2011) 5822-5832.
[2] F. Wanga, L. Chen, C. Deng, H. Ye, X. Jiang, G. Yang, Porous tin film synthesized
by electrodeposition and the electrochemical performance for lithium-ion batteries,
Electrochim. Acta 149 (2014) 330-336.
[3] K. Zhuo, M.-G. Jeong, C.-H. Chung, Highly porous dendritic Ni–Sn anodes for lithium-
ion batteries, J. Power Sources 244 (2013) 601-605.
[4] D.T. Mackay, M.T. Janish, U. Sahaym, P.G. Kotula, K.L. Jungjohann, C.B. Carter, M.G.
Norton, Template-free electrochemical synthesis of tin nanostructures, J. Mater. Sci. 49
(2014) 1476-1483.

13
[5] S. Shin, C. Park, C. Kim, Y. Kim, S. Park, J.-H. Lee, Cyclic voltammetry studies of
copper, tin and zinc electrodeposition in a citrate complex system for CZTS solar cell
application, Curr. Appl. Phys. 16 (2016) 207-210.
[6] M.B. Tahir, Microbial photoelectrochemical cell for improved hydrogen evolution using
nickel ferrite incorporated WO3 under visible light irradiation, Int. J. Hydrogen Energy 44
(2019) 17316-7322.
[7] X. An, S. Li, A. Yoshida, Z. Wang, X. Hao, A. Abudula, G. Guan, Electrodeposition of
Tin-Based Electrocatalysts with Different Surface Tin Species Distributions for
Electrochemical Reduction of CO2 to HCOOH, ACS Sustainable Chem. Eng. 7 (2019) 10
9360-9368.
[8] R.L. Sacci, S. Velardo, L. Xiong, D.A. Lutterman, J. Rosenthal, Copper-Tin Alloys for the
Electrocatalytic Reduction of CO2 in an Imidazolium-Based Non-Aqueous Electrolyte,
Energies 12 (16) (2019) 3132; https://doi.org/10.3390/en12163132.
[9] H.M. Maltanava, T.N. Vorobyova, O.N. Vrublevskaya, Electrodeposition of tin coatings
from ethylene glycol and propylene glycol electrolytes, Surf. Coat. Technol. 254 (2014) 388-
397.
[10] N. Pewnim, S. Roy, Electrodeposition of tin-rich Cu–Sn alloys from a methanesulfonic
acid electrolyte, Electrochim. Acta 90 (2013) 498-506.
[11] N. Kaneko, N. Shinohara, H. Nezu, Effects of aromatic carbonyl compounds on the
surface morphology and crystal orientation of electrodeposited tin from acid stannous sulfate
solutions, Electrochem. Acta 37 (1992) 2403-2409.
[12] Y. Nakamura, N. Kaneko, H. Nezu, Surface morphology and crystal orientation of
electrodeposited tin from acid stannous sulphate solutions containing various additives, J.
Appl. Electrochem. 24 (1994) 569-574.
[13] M. Clarke, S.C. Britton, A New Addition Agent for Bright Tin Plating, Trans. Inst. Metal
finishing 39 (1961) 5-13.
[14] M. Clarke, J.A. Bernie, Abnormal high throwing power and cathode passivity in acid tin
plating baths, Electrochem. Acta 12 (1967) 205-212.
[15] S. Bakkali, T. Jazouli M. Cherkaoui, M. Ebn touhami, N. El Hajjaji, E. Chassaing,
Influence of M12 organic additive on the electrodeposition of tin from an acid sulfate
solution, Plat. Surf. Finish. 90 (2003) 46-49.
[16] F.C. Walsh, C.T.J. Low, A review of developments in the electrodeposition of tin, Surf.
Coat. Technol. 288 (2016) 79-94.
[17] S. Bakkali, A. Benabida, M. Cherkaoui, Tin electrodeposition from sulfate solution
containing a benzimidazolone derivative, Med. J. Chem. 6(2) (2017) 15-22.
14
[18] A.D. Becke, Density-functional exchange-energy approximation with correct asymptotic
behavior, Phys. Rev. A 38 (1988) 3098-3100.
[19] M. El Faydy, B. Lakhrissi, A. Guenbour, S. Kaya, F. Bentiss, I. Warad, A. Zarrouk, In sit
synthesis, electrochemical, surface morphological, UV–visible, DFT and Monte Carlo
simulations of novel 5-substituted-8-hydroxyquinoline for corrosion protection of carbon steel
in a hydrochloric acid solution, J. Mol. Liq. 280 (2019) 341-359.
[20] L.R. Domingo, P. Pérez, J.A. Saez, Understanding the local reactivity in polar organic
reactions through electrophilic and nucleophilic Parr functions, RSC Adv. 3 (2013) 1486-
1494.
[21] S. Kaya, C. Kaya, L. Guo, F. Kandemirli, B. Tüzün, İ. Uğurlu, L.H. Madkour, M.
Saraçoğlu, Quantum chemical and molecular dynamics simulation studies on inhibition
performances of some thiazole and thiadiazole derivatives against corrosion of iron, J. Mol.
Liq. 219 (2016) 497-504.
[22] Y. Arafat, S.T. Sultana, I. Dutta, R. Panat, Effect of Additives on the Microstructure of
Electroplated Tin Films, J. Electrochem. Soc. 165 (16) (2018) D816-D824.
[23] S. Zajkoska, A. Mulone, W. Hansal, U. Klement, R. Mann, W. Kautek, Alkoxylated β-
Naphthol as an Additive for Tin Plating from Chloride and Methane Sulfonic Acid
Electrolytes, Coatings 8(2) (2018) 79; https://doi.org/10.3390/coatings8020079.
[24] C.T.J. Low, F.C. Walsh, Electrodeposition of tin, copper and tin–copper alloys from a
methanesulfonic acid electrolyte containing a perfluorinated cationic surfactant, Surf. Coat.
Technol. 202 (2008) 1339-1349.
[25] Y. Goh, A.S.M.A. Haseeb, M.F.M. Sabri, Effects of hydroquinone and gelatin on the
electrodeposition of Sn–Bi low temperature Pb-free solder, Electrochim. Acta 90 (2013) 265-
273.
[26] C.T.J. Low, F.C. Walsh, The stability of an acidic tin methane sulfonate electrolyte in the
presence of a hydroquinone antioxidant, Electrochim. Acta 53 (2008) 5280-5286.

[27] D. Aranzales, J.H.O.J. Wijenberg, M.T.M. Koper, Voltammetric Study of Tin


Electrodeposition on Polycrystalline Gold from Sulfuric and Methanesulfonic Acid, J.
Electrochem. Soc. 166(8) (2019) D283-D289.
[28] K.K. Yang, M.R. Mahmoudian, M. Ebadi, H.L. Koay, W.J. Basirun, Diffusion
Coefficient of Tin(II) Methanesulfonate in Ionic Liquid and Methane Sulfonic Acid (MSA)
Solvent, Metall. Mater. Trans. B 42(6) (2011) 274-1279.

15
[29] F-.X. Xiao, X-.N. Shen, F-.Z. Ren, A.A. Volinsky, Additive effects on tin
electrodepositing in acid sulfate electrolytes, Int. J. Miner. Metall. Mater. 20(5) (2013) 472-
478.
[30] A. Survila, Z. Mockus, S. Kanapeckaitė, M. Samulevičienė, Effect of sintanol DS-10 and
halides on tin(II) reduction kinetics, Electrochim. Acta 50 (2005) 2879-2885.
[31] C.L. Rinne, J.J. Hren, P.S. Fedkiw, Electrodeposition of Tin Needle-Like Structures, J.
Electrochem. Soc. 149(3) (2002) C150-C158.
[32] S. Bakkali, M. Charrouf, M. Cherkaoui, Effects of additive on the kinetics of tin
electrodeposition from an acidic solution of tin sulfate, Phys. Chem. News 25 (2005) 110-
117.
[33] M. Charrouf, S. Bakkali, M. Cherkaoui, M. EL Amrani, Influence of decyl glucoside on
the electrodeposition of tin, J. Serb. Chem. Soc. 71(6) (2006) 661-668.
[34] J.B. Marro, C.A. Okoro, Y.S. Obeng, K.C. Richardson, The Impact of Organic Additives
on Copper Trench Microstructure, J. Electrochem. Soc. 164(9) (2017) D543-D550.
[35] A. Sharma, S. Bhattacharya, S. Das, K. Das, A study on the effect of pulse
electrodeposition parameters on the morphology of pure tin coatings, Metall. Mater. Trans. A
45 (2014) 4610-4622.
[36] J.-Y. Lee, J.-W. Kim, B-Y. Chang, H.T. Kim, S.-M. Park, Effects of Ethoxylated α-
Naphtholsulfonic Acid on Tin Electroplating at Iron Electrodes, J. Electrochem. Soc. 151(5)
(2004) C333-C341.
[37] C. Meudre, L. Ricq, J.-Y. Hihn, V. Moutarlier, A. Monnin, O. Heintz, Adsorption of
gelatin during electrodeposition of copper and tin–copper alloys from acid sulfate electrolyte,
Surf. Coat. Technol. 252 (2014) 93-101.
[38] P. Zanello, Inorganic Electrochemistry Theory, Practice and Application The Royal
Society of Chemistry 2003.
[39] M.A. Brett and, M.O. Brett, Electrochemistry Principles, Methods and Applications.
Oxford University Press 1994.
[40] X. Ren, M. An, Theoretical and experimental studies of the influence of gold ions and
DMH on cyanide-free gold electrodeposition, RSC Adv. 8 (2018) 2667-2677.
[41] S. Bakkali, R. Touir, M. Cherkaoui, M. Ebn Touhami, Influence of S-
dodecylmercaptobenzimidazole as organic additive on electrodeposition of tin, Surf. Coat.
Technol. 261 (2015) 337-343.
[42] A. Survila, Z. Mockus, S. Kanapeckaite, G. Stalnionis, Kinetics of Sn(II) reduction in
acid sulphate solutions containing gluconic acid, J. Electroanal. Chem. 667 (2012) 59-65.

16
[42] L.A. Azpeitia, C.A. Gervasi, A.E. Bolzán, Electrochemical aspects of tin
electrodeposition on copper in acid solutions, Electrochim. Acta 298 (2019) 400-412.
[44] H. Zarrok, A. Zarrouk, B. Hammouti, R. Salghi, C. Jama, F. Bentiss, Corrosion control
of carbon steel in phosphoric acid by purpald – Weight loss, electrochemical and XPS studies,
Corros. Sci. 64 (2012) 243-252.
[45] N.I. Kairi, J. Kassim, The Effect of Temperature on the Corrosion Inhibition of Mild
Steel in 1 M HCl Solution by Curcuma Longa Extract, Int. J. Electrochem. Sci. 8 (2013)
7138-7155.
[46] M.M. Fares, A.K. Maayta, M.M. Al-Quda, Pectin as promising green corrosion inhibitor
of aluminum in hydrochloric acid solution, Corros. Sci. 60 (2012) 112-117.
[47] G. Chen, J. Lin, Q. Liu, J. Zhang, Y. Wu, H. Li, C. Qu, W. Song, Corrosion inhibition
and the structure-efficiency relationship study of two cationic surfactant, Anti-Corros. Method
M. 66(4) (2019) 388-393.
[48] H.M.A. Soliman, Formalin solution and acetone as organic additives in electrodeposition
of copper, Appl. Surf. Sci. 195 (2002) 155-165.
[49] M.A. El-Okazy, T.M. Zewail, H.A.-M. Farag, Recovery of copper from spent catalyst
using acid leaching followed by electrodeposition on square rotating cylinder, Alex. Eng. J.
57 (2018) 3117-3126.
[50] M. Mobin, M.A. Khan, Investigation on the Adsorption and Corrosion Inhibition
Behavior of Gum Acacia and Synergistic Surfactants Additives on Mild Steel in 0.1 M
H2SO4, J. Dispersion Sci. Technol. 34(11) (2013) 1496-1506.
[51] I. Aiad, S.M. Shaban, H.Y. Moustafa, A. Hamed, Experimental Investigation of Newly
Synthesized Gemini Cationic Surfactants as Corrosion Inhibitors of Mild Steel in 1.0 M HCl,
Prot. Met. Phys. Chem. Surf. 54 (2018) 135-147.
[52] M. Mobin, M. Rizvi, Inhibitory effect of xanthan gum and synergistic surfactant
additives for mild steel corrosion in 1 M HCl, Carbohydr. Polym. 136 (2016) 384-393.
[53] R. Sekar, Synergistic effect of additives on electrodeposition of copper from cyanide-free
electrolytes and its structural and morphological characteristics, Trans. Nonferrous Met. Soc.
China 27 (2017) 1665-1676.

17
[54] P. Dhanasekaran, A. Shukla, K.N. Krishna, I. Rongrin, S.V. Selvaganesh, D. Kalpana,
S.D. Bhat, Enhancing stability and efficiency of oxygen reduction reaction in polymer
electrolyte fuel cells with high surface area mesoporous carbon synthesized from spent
mushroom compost, Sustainable Energy Fuels 3 (2019) 1012-1023.
[55] M.B. Tahir, M. Sagir, M. Zubair, M. Rafique, Ifra Abbas, M. Shakil, I. Khan, S.
Afsheen, A. Hasan, A. Ahmed, WO3 Nanostructures-Based Photocatalyst Approach Towards
Degradation of RhB Dye, J. Inorg. Organomet. Polym. Mater. 28(3) (2018) 1107-1113.
[56] C. Wang, M. An, P. Yang, J. Zhang, Prediction of a new leveler (N-butyl-methyl
piperidinium bromide) for through-hole electroplating using molecular dynamics simulations,
Electrochem. Commun. 18 (2012) 104-107.
[57] C. Wang, J. Zhang, P. Yang, M. An, Electrochemical behaviors of Janus Green B in
through-hole copper electroplating: An insight by experiment and density functional theory
calculation using Safranine T as a comparison, Electrochim. Acta 92 (2013) 356-364.
[58] X. Ren, Y. Song, A. Liu, J. Zhang, G. Yuan, P. Yang, J. Zhang, M. An, D. Matera, G.
Wu, Computational Chemistry and Electrochemical Studies of Adsorption Behavior of
Organic Additives during Gold Deposition in Cyanide-free Electrolytes, Electrochim. Acta
176 (2015) 10-17.
[59] Z. Feng, L. Ren, J. Zhang, P. Yang, M. An, Theoretical Calculations and Electrochemical
Behaviors of Additives in DMH-Based Alkaline Bath for Nanocrystalline Zn-Ni
Electrodeposition, J. Electrochem. Soc. 163(9) (2016) D544-D553.
[60] S.-J. Song, S.-R. Choi, J.-G. Kim, H.-G. Kim, Effect of Molecular Weight of
Polyethylene Glycol on Copper Electrodeposition in the Presence of Bis-3-Sulfopropyl-
Disulfide, Int. J. Electrochem. Sci. 11 (2016) 10067-10079.
[61] C. An, K. An, C. Liu, X. Qu, Adsorption Behavior of 3-phenylacrylaldehyde on a Tin
Surface during Electroplating Int. J. Electrochem. Sci. 12 (2017) 10542-10552.
[62] Z. Lai, S. Wang, C. Wang, Y. Hong, G. Zhou, Y. Chen, W. He, Y. Peng, D. Xiao, A
comparison of typical additives for copper electroplating based on theoretical computation,
Comput. Mater. Sci. 147 (2018) 95-102.
[63] Z. Lai, S. Wang, C. Wang, Y. Hong, Y. Chen, H. Zhang, G. Zhou, W. He, K. Ai, Y.
Peng, Computational analysis and experimental evidence of two typical levelers for acid
copper electroplating, Electrochim. Acta 273 (2018) 318-326.
[64] L. Zheng, W. He, K. Zhu, C. Wang, S. Wang, Y. Hong, Y. Chen, G. Zhou, H.
Miao, J. Zhou, Investigation of poly (1-vinyl imidazole co 1, 4-butanediol diglycidyl ether) as
a leveler for copper electroplating of through-hole, Electrochim. Acta 283 (2018) 560-567.

18
[65] J. Deng, J. Zhang, Y. Tu, P. Yang, M. An, P. Wang, Effect of BEO in the
electrodeposition process of Ni/diamond composite coatings for preparation of ultra-thin
dicing blades: Experiments and theoretical calculations, Ceram. Int. 44(14) (2018) 16828-
16836.
[66] Z. Xiao, Z. Zhou, L. Song, D.X. Wu, C. Zeng, Z. Cao, Electrochemical Studies and
Molecular Dynamics Simulation of the Interaction between Accelerators and Cu Surface
During the Electroplating Process, Int. J. Electrochem. Sci. 14 (2019) 4705-4717.
[67] Z. Feng, L. Wang, D. Li, Q. Sun, P. Xing, M.A.Z. Feng, L. Wang, D. Li, Q. Sun, P.
Xing, M. An, Electrochemical studies of 2-aminopyridine on nanocrystalline Zn-Ni alloy
electrodeposition, J. Electroanal. Chem. 835 (2019) 114-122.
[68] E.A.M. Gad, E.M.S. Azzam, S. Abdel Halim, Theoretical approach for the performance
of 4-mercapto-1-alkylpyridin-1-ium bromide as corrosion inhibitors using DFT, Egypt. J. Pet.
27 (2018) 695-699.
[69] S. Ren, Z. Lei, Z. Wang, Investigation of Nitrogen Heterocyclic Compounds as Levelers
for Electroplating Cu Filling by Electrochemical Method and Quantum Chemical Calculation
J. Electrochem. Soc. 162(10) (2015) D509-D514.
[70] M. Abd El-Raouf, E.A. Khamis, M.T.H. Abou Kana, N.A. Negm, Electrochemical and
quantum chemical evaluation of new bis(coumarins) derivatives as corrosion inhibitors for
carbon steel corrosion in 0.5 M H2SO4, J. Mol. Liq. 255(2018) 341-353.
[71] X. Ren, Y. Song, A. Liu, J. Zhang, G. Yuan, P. Yang, J. Zhang, M. An, D. Matera, G.
Wu, Computational Chemistry and Electrochemical Studies of Adsorption Behavior of
Organic Additives during Gold Deposition in Cyanide-free Electrolytes, Electrochim. Acta
176 (2015) 10-17.
[72] X. Li, S. Deng, H. Fu, T. Li, Adsorption and inhibition effect of 6-benzylaminopurine on
cold rolled steel in 1.0 M HCl, Electrochim. Acta 54 (2009) 4089-4098.
[73] Y.-X. Ji , F.-H. Wang, L.-C. Duan, F. Zhang, X.-D. Gong, Effect of temperature on the
adsorption of sulfanilamide onto aluminum oxide and its molecular dynamics simulations,
Appl. Surf. Sci. 285 (2013) 403-408.
[74] L. Guo, X. Ren, Y. Zhou, S. Xu, Y. Gong, S. Zhang, Theoretical evaluation of the
corrosion inhibition performance of 1,3-thiazole and its amino derivatives, Arab. J. Chem. 10
(2017) 121-130.
[75] J. Xu, B. Chen, J. Lv, D. Chang, D. Niu, S. Hu, X. Zhang, Z. Xin, L. Wang, Aryl
modification of diketopyrrolopyrrole-based quaternary ammonium salts and their applications
in copper electrodeposition, Dyes Pigm. 170 (2019) 107559.

19
[76] R. Wu, X. Qiu, Y. Shi, M. Deng, Molecular dynamics simulation of the atomistic
monolayer structures of N-acylaminoacid-based surfactants, Mol. Simul. 43 (2017) 491-501.
[77] S.-W. Xie, Z. Liu, G.-C Han, W. Li, J. Liu, Z. Chen, Molecular dynamics simulation of
inhibition mechanism of 3,5-dibromo salicylaldehyde Schiff’s base, Comput. Theor. Chem.
1063 (2015) 50-62.
[78] M. El Faydy, F. Benhiba, B. Lakhrissi, M. Ebn Touhami, I. Warad, F. Bentiss, A.
Zarrouk, The inhibitive impact of both kinds of 5-isothiocyanatomethyl-8- hydroxyquinoline
derivatives on the corrosion of carbon steel in acidic electrolyte, J. Mol. Liq. 295 (2019)
111629.
Figures

Fig. 1. Optimized molecular structure of decyl glucoside.

0.00
Current density (A/cm2)

-0.05

-0.10 a
b
c
-0.15

-0.20

-0.25
-1.2 -1.0 -0.8 -0.6 -0.4
Potential (V/SCE)

Fig. 2. Linear sweep voltammogram on iron working electrode at 20 mV S-1: a) 10-3M


additive; b) 0.056 M H2 SO4; c) 10-3 M additive + 0.056 M H2 SO4.

20
Fig. 3. Current-voltagecurves obtained at 5 mV s-1 in solutions: a) in the absence of additive;
b) in the presence of additive.

0.002
Current density (A/cm2)

0.000

-0.002

-0.004

050 mV/s
-0.006 100 mV/s
150 mV/s
200 mV/s
-0.008 250 mV/s

-0.010
-0.8 -0.7 -0.6 -0.5 -0.4
Potential (V/SCE)

Fig. 4. Linear sweep voltammograms for Tin reduction in absence of additive at various
scanning rates.

21
0.000

Current density (A/cm2)


-0.001

-0.002

050 mV/s
-0.003 100 mV/s
150 mV/s
200 mV/s
250 mV/s
-0.004

-0.8 -0.7 -0.6 -0.5 -0.4


Potential (V/SCE)

Fig. 5. Linear sweep voltammograms for Tin reduction in presence of additive at various
scanning rates.

Fig. 6. The curve Epc versus the logarithm of the scan rate: a) in the absence of additive; b) in
the presence of additive.

22
0.008

0.007

0.006
Ip (A/cm2)

a
b
0.005

0.004

0.003

0.002

0.2 0.3 0.4 0.5

(u (rad/s)0.5

Fig. 7. The variation of the absolute value of I pvs.ν 1/2 rate: a) in the absence of additive; b) in
the presence of additive.

0.6
Surface coverage ()

0.5

0.4

0.3
295 300 305 310 315 320 325 330

Temperature (K)

Fig. 8. Variation of the surface coverage with the temperature.

23
-2.3
a (r2 = 0.9789)
-2.4 b (r2 = 0.9949)

-2.5

Log [I(A)]
-2.6

-2.7

-2.8

-2.9
0.0030 0.0031 0.0032 0.0033 0.0034
1/T(K-1)

Fig. 9. Relationship between current density at E = -610 mV/SCE and T−1: a) in the absence
of additive; b) in the presence of additive

Fig. 10. SEM image of deposition: a) in the absence of additive; b) in the presence of additive
(e = 10 μm, I = 10 mA/cm2and  = 2000 tr mn-1).

1.0
(112)

b
0.8

0.6
(101)

0.4

0.2
I/Imax

0.0
1.0
a
(220)

0.8

0.6
(200)

0.4
(211)

(400)
(301)
(101)

(321)
(112)

0.2

0.0
30 35 40 45 50 55 60 65

Fig. 11. XRD patterns of electroplated Sn film: a) in the absence of additive; b) in the presence of
additive.

24
(1 1 2)
6 b

(3 0 1)
Texture coefficient Tc

(2 2 0)
(1 0 1)
2

0
-- -- --

(2 2 0)
6 a

(4 0 0)
(3 0 1)
(2 0 0)

(2 1 1)

(3 2 1)
2

(1 1 2)
(1 0 1)

0
----

Planes
Fig. 12. Texture coefficient of electroplated Sn film: a) in the absence of additive; b) in the
presence of additive.
HOMO LUMO

Fig. 13. Schematic representation of HOMO and LUMO molecular orbital of the studied
molecule.

Fig. 14. Electrostatic potential (ESP) maps for additive molecule.

25
Fig. 15. a) Temperature and b) energy fluctuations in the interaction process of the additive on
Fe (1 1 0) surface.

26
Fig. 16.The adsorption processes of the additive at 298 K.

7 o
4.48 A
6

4
g (r)

0 5 10 15 20
o
r (A)

Fig. 17. Radial distribution functionof the additive on Fe(1 1 0) surface in water solution.
Tables
Table 1
Electrolyte composition
Electrolyte SnSO4 H2SO4 Additive
(mol L-1) (mol L-1) (mol L-1)
With additive 0.014 0.06 0
Without additive 0.014 0.06 10-3

Table 2
αc values in the absence and presence of the additive
Without Additive With Additive
2
Correlation coefficient (R ) 0.97968 0.95853
αc 0.15 0.16

Table 3
Effect of the additive on the diffusion coefficient of the stannous ion
Without Additive With Additive
Correlation coefficient (R 2 ) 0.98602 0.97475
2 -6
D0 (cm /S) 5.6×10 1.18×10-6

Table 4
Activation energy values for each electrolyte
Without Additive With Additive
2
Correlation coefficient (R ) 0.96644 0.99782
Ea (kJ/mol) 12.42 23.61
27
Table 5
Calculated quantum chemical parameters for additive molecule in aqueous phase
Quantum chemical E ELUMO E   
HOMO
parameters (eV) (Debye) (eV) (eV)
(eV) (eV)
Value -7.42 -0.30 7.12 5.49 3.86 3.56

Table 6
Theoretical prediction of reactive sites using Parr function for the additive molecule
Atom Pk Pk Nk k
1 C 0.093 --------- 0.156 ---------
2 C 0.025 0.6560 0.043 1.3810
3 C 0.031 0.0100 0.052 0.0270
5 C 0.030 0,0770 0.051 0.1620
11 O 0.356 0.0020 0.600 0.0004
12 O 0.016 0.0002 0.026 ---------
13 O 0.063 0.0008 0.105 0.0018
15 O 0.108 --------- 0.178 ---------
17 O 0.156 --------- 0.262 ---------
53 O 0.030 0.0070 0.05 0.0160

Conflict of Interest

The authors have declared no conflict of interest

28

You might also like