Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Science and Engineering C 63 (2016) 376–383

Contents lists available at ScienceDirect

Materials Science and Engineering C

journal homepage: www.elsevier.com/locate/msec

Effect of silver on microstructure and antibacterial property of 2205


duplex stainless steel
Sheng-Min Yang a, Yi-Chun Chen a, Yeong-Tsuen Pan b, Dong-Yih Lin a
a
Department of Chemical and Materials Engineering, National University of Kaohsiung, 700, Kaohsiung University Rd., Nanzih District, Kaohsiung 811, Taiwan, ROC
b
New Materials R&D Department, China Steel Corporation, 1, Chung Kang Rd., Hsiao Kang, Kaohsiung 81233, Taiwan, Taiwan, ROC

a r t i c l e i n f o a b s t r a c t

Article history: In this study, 2205 duplex stainless steel (DSS) was employed to enhance the antibacterial properties of material
Received 20 October 2015 through silver doping. The results demonstrated that silver-doped 2205 DSS produces an excellent bacteria-
Received in revised form 1 February 2016 inhibiting effect against Escherichia coli and Staphylococcus aureus. The antibacterial rates were 100% and 99.5%,
Accepted 2 March 2016
respectively. Because the mutual solubility of silver and iron is very low in both the solid and liquid states, a
Available online 3 March 2016
silver-rich compound solidified and dispersed at the ferrite/austenite interface and the ferrite, austenite, and sec-
Keywords:
ondary austenite phases in silver-doped 2205 DSS. Doping 2205 DSS with silver caused the Creq/Nieq ratio of fer-
2205 duplex stainless steel (DSS) rite to decrease; however, the lower Creq/Nieq ratio promoted the rapid nucleation of γ2-austenite from primary
Microstructure α-ferrite. After 12 h of homogenisation treatment at 1200 °C, the solubility of silver in the γ-austenite and α-
Phase transformation ferrite phases can be increased by 0.10% and 0.09%, respectively. Moreover, silver doping was found to accelerate
Homogenizing the dissolution of secondary austenite in a ferrite matrix during homogenisation.
Silver © 2016 Published by Elsevier B.V.

1. Introduction et al. [12] reports, bacteria are a natural part of the raw materials for
food production but can also reside in food-processing equipment
Bacterial infection is a serious issue among fields in which critical stan- where they can (re)contaminate food products. Ready to-eat foods are
dards of sanitation must be maintained, such as the food processing and potentially high risk products since they do not receive further heat
health care industries. Especially for hospital-acquired infections, bacteri- treatment before consumption. This makes them vulnerable to (re)con-
al spreading often occurs through contact with medical appliances. Anti- tamination and, if storage conditions allow further growth of bacteria,
bacterial materials have been developed for countering this problem, and they may cause disease in humans [13]. Therefore, reducing bacteria re-
stainless steel with biological properties has been widely used. side in food-processing equipment may prevent adhering to surfaces of
Both silver and copper are antibacterial elements used in fabricating food production by silver-containing stainless steel for antibacterial ap-
stainless steel. To obtain antibacterial properties through copper dop- plications in food-processing industry. Up to now, the relative litera-
ing, Hong and Koo [1] administered an aging treatment to acquire a pre- tures of silver-containing duplex stainless steel (DSS) in antibacterial
cipitate concentration of copper ions on the surface, as is the standard application is few to publish. This considered, the purpose of this
procedure for manufacture. However, doping stainless steel with silver study was to investigate the effect of silver on the microstructure and
can produce a superior antibacterial effect without aging treatment. antibacterial rate by doping a 2205 DSS matrix with silver through melt-
Therefore, the application of Ag-doped steel has been extended to ing and thus determine whether this process yields enduring antibacte-
kitchens and food processing factories [2,3]. Many methods have been rial properties.
developed and adopted to produce antibacterial film on the surfaces
of materials, such as ion implantation [4], chemical assembly [3], and
plasma deposition [5,6]. These developments have been made in re- 2. Experimental
sponse to an observable loss of efficacy of antibacterial properties,
once the thin antibacterial surface layer is damaged. 2.1. Material preparation
Duplex stainless steel (DSS) was developed to function in aggressive
environments. It consists of an approximately equal proportion of an 2205 DSS was doped with 0.2 wt.% silver in this study. The chemical
austenite phase and ferrite phase. This phase constitution provides ex- compositions of the steel are listed in Table 1, and the specimens were
cellent properties of hardness, toughness, and localised corrosion resis- named 2205-Base and 2205–0.2 Ag. The materials were melted in a
tance [7–10]. It is the most widely used in pulp and paper industry, high-frequency induction furnace under a nitrogen atmosphere and
cargo tanks for ships and trucks, food processing equipment, and homogenised at 1120 °C for 2 h to eliminate microsegregation after
biofuels plants. According to Michiels et al. [11] and Fonnesbech Vogel casting.

http://dx.doi.org/10.1016/j.msec.2016.03.014
0928-4931/© 2016 Published by Elsevier B.V.
S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383 377

Table 1 2.3. Homogenisation treatment


Chemical composition of 2205-Base and 2205–0.2 Ag duplex stainless steel unit: wt.%.

Specimens C Si Mn P S Cr Ni Mo Ag N Fe To investigate the microstructures of 2205-Base and 2205–0.2 Ag


2205-Base 0.02 0.38 1.19 0.023 0.001 23.4 6.2 3.18 – 0.22 Bal.
DSS, the specimens were placed in a box furnace for secondary heat
2205–0.2 0.02 0.38 0.81 0.024 0.003 22.8 6.5 2.97 0.19 0.26 Bal. treatment at 1200 °C. During homogenisation, the specimens were
Ag water quenched after 2, 4, 6, 8, 10, and 12 h treatment to prevent the
reprecipitation of second phases.

2.4. Phase quantitation and microstructure examination

To examine the effect of silver doping on 2205 DSS, the Fischer


Feritscope FMP30 was employed to measure variation in the δ-ferrite
content after various homogenisation durations. The Feritscope
operates by generating an alternating magnetic field in the sample
which is proportional to the ferrite content and which the instrument
Fig. 1. Schematics of the film fabrication method. can detect. As in metallography preparation, the 2205-Base and 2205–
0.2 Ag specimens were ground using SiC abrasive papers to the #2000
grade and polished with 1.0- and 0.3-μm Al2O3 powder. An LB1 reagent
2.2. Antibacterial property examination
solution (0.5 g of K2S2O5 + 20 g of NH4HF2 + 100 mL of H2O) was then
used to etch phases. X-ray diffractometer (XRD; Bruker D8) was
The antibacterial procedure was carried out according to JIS Z2801:
adopted to detect the structures with copper target and scanning rate
2000 specifications. The solution-treated specimens were cut into
of 0.05°. In addition, optical microscopy and field emission scanning
squares with dimensions of 50 mm × 50 mm × 5 mm. The antibacterial
electron microscopy (FE-SEM; Hitachi S-4800) with energy-dispersive
specimens were polished, cleaned with acetone using an ultrasonic ma-
spectroscopy (EDS) were employed to observe and analyse the micro-
chine, and dried thoroughly. The test was performed using bacteria-
structures. An electron probe X-ray microanalyser (JEOL JXA-8200)
containing solutions (suspensions) held in close contact with test sur-
was also employed to examine the silver content by using
faces. Escherichia coli ATCC6538P and Staphylococcus aureus
wavelength-dispersive spectroscopy (WDS). In order to characterise
ATCC6538P were used as test organisms in this study. E. coli and
the surface morphology of the 2205 DSS, atomic force microscopy
S. aureus were cultivated at 37 °C in a nutrient broth solution. The initial
(AFM; NT-MDT & NTFGRA Spectra-upright) was performed.
bacterial concentration was approximately 105 CFU mL−1, and 0.4 mL of
the bacterial suspension was then applied dropwise and spread over
3. Results and discussion
each sample to create a contaminated surface. The samples were subse-
quently covered by a sterilised polyethylene film (4 cm × 4 cm) to keep
3.1. Effect of silver doping on microstructure of 2205 duplex stainless steel
them in close contact with the bacterial suspension and incubated at
37 °C for 24 h. After incubation, 10 mL of the diluted bacterial suspen-
Phase identification is confirmed by X-ray diffraction patterns, as
sion was pipetted onto nutrition agar plates to collect live bacteria. A
shown in Fig. 2. The major X-ray diffraction peaks are composed of fer-
schematic of the film fabrication method is shown in Fig. 1. The antibac-
rite and austenite. Due to low content of 0.2 wt.% in 2205–0.2 Ag DSS
terial rates can be calculated using the following formula:
steel, the intensity is weak to detect the silver diffraction peak. In addi-
N1‐N2 tion, γ (111) and α (110) peaks are the preferred orientation in 2205-
Antibacterial rateð%Þ ¼  100 ð1Þ Base and 2205–0.2Ag DSS steel, respectively. This indicates that the
N1
major diffraction peak of ferrite increased by doping of silver.
The typical microstructure of 2205 DSS after solution treatment,
consisting of α-ferrite, primary γ-austenite (γ1), and secondary γ-
N1: number of viable bacteria adhered to specimens after 24 h of austenite (γ2), is shown in Fig. 3. As described by Knyazeva et al. [10]
incubation. and Rajasekhar et al. [14], the quasibinary section of the Fe\\Cr\\Ni
N2: number of bacteria adhered to specimens after 24 h of incubation. phase diagram at 70% iron shows that 2205 DSS undergoes dual-phase

Fig. 2. X-ray diffraction diagrams of the duplex stainless steel: (a) 2205-Base; and (b) 2205–0.2 Ag.
378 S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383

Fig. 3. Optical micrographs of the duplex stainless steel: (a) 2205-Base; and (b) 2205–0.2 Ag.

solidification in the F mode (i.e., L → L + α → α + γ). Fig. 3 also shows 0.77 ± 0.26 wt.% and 0.65 ± 0.25 wt.%. Knyazeva and Pohl [10] reported
that γ2 precipitated from the ferrite phase in both the 2205-Base and that the nitrogen solubility of ferritic-austenitic steel could be increased
2205–0.2 Ag specimens. Knyazeva et al. [10] and Shek et al. [15] stated by increasing the chromium content. Consistent with these findings, the
that γ2 can be classified according to phase transformation behaviour, nitrogen content of γ2-austenite in 2205–0.2 Ag DSS increased with an
which is produced by solid solution transformation from the ferrite increasing chromium content. While the higher nitrogen content in-
phase and ferrite decomposed by a eutectoid solid-state reaction creased the Nieq value, silver doping caused a decrease in the Creq
(α → σ + γ2). Because the 2205-Base and 2205–0.2 Ag specimens value, causing an overall decrease in the Creq/Nieq ratio in 2205–0.2 Ag
were solute treated and then subjected to rapid quenching, γ2 formed DSS. This suggests that the lower Creq/Nieq ratio of ferrite promoted
from the ferrite phase without σ phase precipitation. This indicates the rapid nucleation of γ2-austenite from α1-ferrite, competing with
that the γ2-austenite decomposed into ferrite phase. In addition, the grain growth. The Creq/Nieq ratio of α2-ferrite in 2205–0.2 Ag DSS was
grain size of γ2 in 2205–0.2 Ag DSS was smaller than that in 2205- lower than that in 2205-Base DSS, as shown in Tables 2 and 3. Finer
Base DSS with a large amount of nuclei. grains of γ2-austenite were also observed in 2205–0.2 Ag DSS, as
Hammer and Svensson's [16] equation was adopted to examine the shown in Fig. 3.
Creq/Nieq ratios of α1, α2, γ1, and γ2 phases in various locations. SEM- In general, the ferritic dendrite precipitated and grew in 2205 DSS
EDS was performed to identify the chemical composition of selected during initial solidification after melting. However, the ferritic dendrite
phases, with the results listed in Table 2 and Tables 3. These results was subsequently interrupted by the solidified austenitic dendrite.
showed that silver doping resulted in a decrease in the Creq/Nieq ratios Flemings [17] noted that alloys or impurity elements, known as second-
of α1-ferrite and α2-ferrite in 2205–0.2 Ag DSS. The Creq/Nieq ratio of ary inclusions, were usually rejected to interdendritic spaces before
γ2-austenite was additionally observed to be lower than that in 2205- complete solidification. Secondary inclusions are usually smaller than
Base DSS. 2205–0.2 Ag DSS was found to possess a higher nitrogen con- dendrite arm spacing, measuring 0.1 to 5 μm for ferrous casting and in-
tent in γ2-austenite than 2205-Base DSS did, with respective values of gots. However, 2205 DSS transforms completely into austenite at

Table 2
Chemical composition and Creq/Nieq ratio of α-ferrite and γ-austenite on different locations in 2205-Base DSS.*

2205-Base N Si Cr Mn Ni Mo Creq/Nieq ratio

α1 0.43 ± 0.20 0.54 ± 0.18 22.97 ± 0.85 0.94 ± 0.25 4.86 ± 0.32 5.65 ± 0.51 2.79
γ1 0.91 ± 0.23 0.46 ± 0.06 20.62 ± 0.67 1.31 ± 0.39 6.42 ± 0.47 2.94 ± 0.31 1.28
α2 0.43 ± 0.32 0.54 ± 0.13 23.27 ± 0.55 1.08 ± 0.42 5.07 ± 0.21 5.84 ± 0.70 2.78
γ2 0.65 ± 0.25 0.38 ± 0.08 20.35 ± 0.43 1.56 ± 0.20 6.84 ± 0.52 2.70 ± 0.30 1.48
*
α grain into γ2 grows was termed as “α2” and the α grain without γ phase as “α1”.

Table 3
Chemical composition and Creq/Nieq ratio of α-ferrite and γ-austenite on different locations in 2205–0.2 Ag DSS.*

2205–0.2Ag N Si Cr Mn Ni Mo Creq/Nieq ratio

α1 0.64 ± 0.14 0.54 ± 0.05 23.53 ± 0.48 1.31 ± 0.24 5.30 ± 0.08 5.03 ± 0.21 2.24
γ1 0.84 ± 0.22 0.46 ± 0.09 20.40 ± 0.46 1.53 ± 0.31 6.84 ± 0.50 2.63 ± 0.27 1.28
α2 0.64 ± 0.17 0.63 ± 0.15 22.38 ± 0.79 1.11 ± 0.27 5.27 ± 0.55 5.77 ± 0.43 2.12
γ2 0.77 ± 0.26 0.35 ± 0.07 20.16 ± 0.69 1.31 ± 0.31 6.83 ± 0.60 2.77 ± 0.57 1.34
⁎ α grain into γ2 grows was termed as “α2” and the α grain without γ phase as “α1”.
S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383 379

Fig. 5. Three-dimensional representation of AFM image showing the Ag particles


distribution in 2205–0.2 Ag duplex stainless steel.

silver solidified and caused silver particles to disperse in 2205–0.2 Ag


DSS, as shown in Fig. 4 (a). Silver particles were located at the ferrite
phase, the austenite phase, and the ferrite/austenite interface. A quanti-
tative method was used to measure the position of silver particles in a
10-mm2 area; the results showed that the probability of silver particles
forming at the ferrite/austenite interface was higher than that in other
locations, as shown in Fig. 4 (b). Swartzendruber [18] reported that
the mutual solubility of Ag and Fe is very low in both the solid and liquid
states. The solubility of Ag in solid Fe reaches a maximum of approxi-
mately 0.022 at.% at 1398 °C in the γ-phase. This indicates that silver
particles tend to solidify at interdendritic spaces because of secondary
inclusions with low melting points. Fig. 4(c) shows statistics of correla-
tion between quantity and Ag particle size distribution in 2205–0.2 Ag
duplex stainless steel. The distribution of silver particles size is from sev-
eral nanometres to 7 μm, but more than 50% of the silver particle size is
between 0.5 and 1.5 μm. The morphology of silver particles was then
characterized by the atomic force microscope (AFM), which reveals a
presence of metallic particles with different sizes, as shown in Fig. 5.
The average roughness (Sa) and Root Mean Square (Sq) of 2205–0.2
Ag duplex stainless steel are 10.10 ± 0.2 nm and 13.47 ± 1.6 nm,
respectively.

3.2. Antibacterial performance

Tables 4 and 5 show results regarding the bacteria-inhibiting effects


of the 2205-Base and 2205–0.2 Ag specimens. The 2205-Base DSS

Table 4
The antimicrobial rate of E. coli for 2205-Base and 2205–0.2 Ag duplex stainless steel.

Specimen ID Bacteria count/CFU/mL Antibacterial rate (%)

Initial 24 h 24 h

2205-Base 1.1 × 105 2.7 × 106 –


Fig. 4. (a) Backscattered SEM image of 2205–0.2 Ag duplex stainless steel; (b) statistics of 2205–0.2 Ag 1.1 × 105 0 100.00
correlation between quantity and Ag particle location in 2205–0.2Ag duplex stainless
steel; (c) statistics of correlation between quantity and Ag particle size distribution in
2205–0.2Ag duplex stainless steel.

Table 5
The antimicrobial rate of S.·aureus for 2205-Base and 2205–0.2 Ag duplex stainless steel.
temperatures below approximately 1100 °C. This indicates the potential Specimen ID Bacteria count/CFU/ml Antibacterial rate (%)
for the composition-invariant transformation of ferrite into austenite
Initial 24 h 24 h
below the solvus [10]. In addition, silver possesses a low melting point
of 961 °C, facilitating the formation of liquid silver in the microstructure 2205-Base 2.3 × 105 3.8 × 106 –
2205–0.2 Ag 2.3 × 105 1.0 × 103 99.56
during the homogenisation process at 1120 °C. Subsequently, liquid
380 S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383

Fig. 6. Mapping of silver distribution in 2205–0.2 Ag DSS after homogenizing at 1200 °C for 12 h.

produced no antibacterial effect in inhibiting E. coli and S. aureus, yield- solubility of silver increased in both the α and γ phases, which
ing an antibacterial rate of zero. The 2205–0.2 Ag DSS exhibited an ex- contained silver at 0.090 ± 0.010 and 0.100 ± 0.020 wt.%, respec-
cellent bacteria-inhibiting effect on E. coli, with an antibacterial rate of tively, after homogenisation treatment. Liao et al. [19] demonstrated
100%, and produced a similar effect on S. aureus, with an antibacterial that silver particles could form an Ag-rich compound, because silver
rate of N 99.5%. Liao et al. [19] found that the antibacterial rate of AISI may bind with other alloys. The silver solidified at the end of the
304 austenitic stainless steel was considerably lower, at 0.2%; however, casting process, resulting in a reaction between the silver and other
steel with an Ag content of 0.3 wt.% was found to possess excellent an- alloys through diffusion to form an Ag-rich compound. The chemical
tibacterial properties against E. coli. composition of the Ag-rich compound was identified, consisting of
It is well known that the antibacterial activity of silver is depen- silver at N 88 wt.% and other alloys including Fe, Cr, O, and S. The ef-
dent on the silver cation Ag + , which binds strongly to electron fect of homogenisation on the composition of the Ag-rich compound
donor groups in biological molecules containing sulphur, oxygen or was not obvious in this study.
nitrogen. On the other hand, Ag can react with amino acids Knyazeva and Pohl stated that the ferrite-austenite relationship in
(H2NCHRCOOH, where R is an organic substituent) or amino groups 2205 DSS depends on the specific heat treatment and metallurgy; how-
(\\NH2) of membranes or enzymes inside bacterial cells, which ever, the phase fractions are usually between 40:60 and 60:40. In this
eventually kills bacteria [20,21]. Even though the 2205–0.2 Ag DSS study, 2205-Base and 2205–0.2 Ag DSS consisted of 51.8% and 52.9% fer-
releases small amount of silver ions in bacteria-containing solutions rite, respectively. This indicates that adding 0.2% silver to 2205 DSS in-
during the antibacterial procedure, Ag concentration is not high creased the ferrite content by 1.1%. Fig. 7 shows that the ferrite
enough to obtain significant bacterial inhibition. And therefore content continually decreased in both 2205-Base and 2205–0.2 Ag DSS
when bacteria are present and produce some chemical compounds until a homogenisation time of 6 h, at which point the ferrite contents
through their metabolism, the effect of increased Ag dissolution were 50.5% and 49.2% respectively. However, ferrite content increased
can occur and furthermore improve the bacterial inhibition of the gradually between 8 and 12 h, with both DSSs possessing a ferrite con-
2205–0.2 Ag DSS surfaces [22]. tent of approximately 51.8%. This shows that the homogenisation time

3.3. Solubility of silver and microstructure characterisation during homog-


enisation treatment

After homogenisation at 1200 °C for 12 h, the silver particles ex-


hibited a spherical morphology, dispersing at the ferrite/austenite
interface and the austenitic and ferritic phases, as shown in Fig. 6.
A silver-iron phase diagram showed that the solubility of Fe in Ag
is between 4 and 6 ppm by weight in the temperature range of
1000–1600 °C [18]. Table 6 shows the values for the solubility of sil-
ver in the α and γ phases, as identified using EPMA-WDS. The

Table 6
Solubility of silver on γ-austenite and α-ferrite phases in 2205–0.2 Ag duplex stainless
steel.

2205–0.2 Ag DSS

non-homogenizing 1200 °C, 12 h

γ 0.008 ± 0.004 0.100 ± 0.020


α 0.009 ± 0.004 0.090 ± 0.010
Fig. 7. The relationship of ferrite content and different homogenizing times at 1200 °C.
S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383 381

Fig. 8. Microstructure of 2205-Base DSS homogenizing at 1200 °C for 2–12 h.

of 6 h was a critical point in the phase transformation of ferrite into aus- whereby the dissolution of the γ2 phase can be accelerated in a ferrite
tenite. The austenite phase transformed into ferrite from 6 to 12 h dur- matrix by adding a silver alloy.
ing homogenisation. Regarding high-temperature diffusion, the
diffusion rate of chromium and nickel was approximately 100 times 4. Conclusions
higher in ferrite than in austenite. Lippold et al. [23] and Sieurin and
Sandström [24] found that the transformation of ferrite into austenite In this study, the antibacterial properties of 2205 DSS were observed
is a predominant mechanism during phase transformation at 1200 °C. after the addition of 0.2 wt.% silver. Because the mutual solubility of Ag
Moreover, there was a notable variation in the ferrite content due to sil- and Fe is very low in both the solid and liquid states, the silver solidified
ver doping in 2205 DSS. and reacted with impurities to form a silver-rich compound during so-
Figs. 8 and 9 show the microstructures of 2205-Base and 2205–0.2 lidification. The probability of silver-rich compound distribution at the
Ag DSS. The γ2 phase can be observed in 2205-Base DSS homogenised ferrite/austenite interface was higher than that in the ferrite, austenite,
at 1200 °C for 2 and 4 h; however, the γ2 phase clearly dissolved in a fer- and secondary austenite phases. In addition, the lower Creq/Nieq ratio of
rite matrix at 6 h. In addition, the γ2 phase evidently dissolved after 4 h ferrite promoted the rapid nucleation of secondary austenite from α-
of homogenisation in 2205–0.2 Ag DSS. Other precipitates, such as car- ferrite through the addition of silver. After homogenisation treatment,
bide, nitride, the χ-phase, and the σ-phase, were not observed in either the solubility of silver on the γ-austenite and α-ferrite phases was
steel because they cannot form or exist at 1200 °C. Knyazeva and Pohl 0.10% and 0.09% respectively in silver-doped 2205 DDS. The results
[19,25], Kashiwar et al. [26] Lee et al. [27], Chen et. al. [28], Gregori et. showed the antibacterial rates of silver-doped 2205 DSS against E. coli
al. [29], and Yang et al. [30] have reported a relative phenomenon and S. aureus were 100% and 99.5%, respectively.
382 S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383

Fig. 9. Microstructure of 2205–0.2Ag DSS homogenizing at 1200 °C for 2–12 h.

Acknowledgement [5] K. Jamuna-Thevi, S.A. Bakar, S. Ibrahim, N. Shahab, M.R.M. Toff, Quantification of sil-
ver ion release, in vitro cytotoxicity and antibacterial properties of nanostuctured Ag
doped TiO2 coatings on stainless steel deposited by RF magnetron sputtering, Vacu-
This study was funded in part by the National Science Council of um 86 (2011) 235–241.
Taiwan under Grant NSC 101-2811-E-390-001. [6] S. Zanna, C. Saulou, M. Mercier-Bonin, B. Despax, P. Raynaud, A. Seyeux, P. Marcus,
Ageing of plasma-mediated coatings with embedded silver nanoparticles on stain-
less steel: an XPS and ToF-SIMS investigation, Appl. Surf. Sci. 256 (2010) 6499–6505.
[7] J. Olsson, M. Snis, Duplex–a new generation of stainless steels for desalination
References plants, Desalination 205 (2007) 104–113.
[8] V. Muthupandi, Bala P. Srinivasan, S.K. Seshadri, S. Sundaresan, Effect of weld metal
[1] I.T. Hong, C.H. Koo, Antibacterial properties, corrosion resistance and mechanical chemistry and heat input on the structure and properties of duplex stainless steel
properties of Cu-modified SUS 304 stainless steel, Mater. Sci. Eng. A393 (2005) welds, Mater. Sci. Eng. A 358 (2003) 9–16.
213–222. [9] E.F. Monlevade, I.G.S. Falleiros, Orientation relationships associated with austenite
[2] M. Hjelm, L.R. Hilbert, P. Møller, L. Gram, Comparison of adhesion of the food spoil- formation from ferrite in a coarse grained duplex stainless steel, Metall. Mater.
age bacterium Shewanella putrefaciens to stainless steel and silver surfaces, J. Appl. Trans. A 37 (2006) 939–949.
Microbiol. 92 (2002) 903–911. [10] M. Knyazeva, M. Pohl, Duplex steels: part I: genesis, formation, structure, metallog-
[3] L. Chen, L. Zheng, Y. Lv, H. Liu, G. Wang, N. Ren, D. Liu, J. Wang, Robert I. Boughton, raphy, Metallogr. Microstruct. Anal. 2 (2013) 113–121.
Chemical assembly of silver nanoparticles on stainless steel for antimicrobial appli- [11] C.W. Michiels, M. Schellekens, C.C.F. Soontjens, K.J.A. Hauben, Molecular and meta-
cations, Surf. Coat. Technol. 204 (2010) 3871–3875. bolic typing of resident and transient fluorescent pseudomonad flora from a meat
[4] J. Dudognon, M. Vayer, A. Pineau, R. Erre, Mo and Ag ion implantation in austenitic, mincer, J. Food Prot. 60 (1997) 1515–1519.
ferritic and duplex stainless steels: a comparative study, Surf. Coat. Technol. 203 [12] B. Fonnesbech Vogel, H.H. Huss, B. Ojeniyi, P. Ahrens, L. Gram, Appl. Environ.
(2008) 180–185. Microbiol. 60 (2001) 2586–2595.
S.-M. Yang et al. / Materials Science and Engineering C 63 (2016) 376–383 383

[13] H. Ericsson, A. Eklöw, M.-L. Danielsson-Tham, S. Loncarevic, L.-O. Mentzing, I. [23] John C. Lippold, Damian J. Kotecki, Welding Metallurgy and Weldability of Stainless
Persson, H. Unnerstad, W. Tham, J. Clin. Microbiol. 35 (1997) 2904–2907. Steels, 1 ed. Wiley-Interscience, New Castle, 2005.
[14] K. Rajasekhar, C.S. Harendranath, R. Raman, S.D. Kulkarni, Microstructural evolution [24] H. Sieurin, R. Sandström, Austenite reformation in the heat-affected zone of duplex
during solidification of austenitic stainless steel weld metals: a color metallographic stainless steel 2205, Mater. Sci. Eng. A418 (2006) 250–256.
and electron microprobe analysis study, Mater. Charact. 38 (1997) 53–65. [25] M. Knyazeva, M. Pohl, Duplex steels: part II: carbides and nitrides, Metallogr.
[15] C.H. Shek, C. Dong, J.K.L. Lai, K.W. Wong, Early-stage Widmanstätten growth of the γ Microstruct. Anal. 2 (2013) 343–351.
phase in a duplex steel, Metall. Mater. Trans. A 31 (2000) 15–19. [26] A. Kashiwar, N. Phani Vennela, S.L. Kamath, R.K. Khatirkar, Effect of solution anneal-
[16] O. Hammer, U. Svensson, Solidification Technology in the Foundry and Casthouse, ing temperature on precipitation in 2205 duplex stainless steel, Mater. Charact. 74
Metals Society, London, 1983. (2012) 55–63.
[17] Merton C. Flemings, Solidification Processing, McGraw-Hill, New York, 1974. [27] K.M. Lee, H.S. Cho, D.C. Choi, Effect of isothermal treatment of SAF 2205 duplex
[18] S.L.J. wartzendruber, The Ag-Fe (Silver-Iron) system, Bull. Alloy Phase Diagr. 5 stainless steel on migration of δ/γ interface boundary and growth of austenite, J. Al-
(1984) 560–564. loys Compd. 285 (1999) 156–161.
[19] K.-H. Liao, K.-L. Ou, H.-C. Cheng, C.-T. Lin, P.-W.P. eng, Effect of silver on antibacterial [28] T.H. Chen, J.R. Yang, Effect of solution treatment and continuous cooling on σ-phase
properties of stainless steel, Appl. Surf. Sci. 256 (2010) 3642–3646. precipitation in a 2205 duplex stainless steel, Mater. Sci. Eng. A311 (2001) 28–41.
[20] S.L. Percival, P.G. Bowler, D. Russell, Bacterial resistance to silver in wound care, J. [29] A. Gregori, J.-O. Nilsson, Decomposition of ferrite in commercial superduplex stain-
Hosp. Infect. 60 (2005) 1–7. less steel weld metals: microstructural transformations above 700 °C, Metall. Mater.
[21] N.C. Kasuga, M. Sato, A. Amano, A. Hara, S. Tsuruta, A. Sugie, K. Nomiya, Light-stable Trans. 33A (2002) 1009–1018.
and antimicrobial active silver(I) complexes composed of triphenylphosphine and [30] S.-M. Yang, Y.-C. Chen, C.-H. Chen, W.-P. Huang, D.-Y. Lin, Microstructural character-
amino acid ligands: synthesis, crystal structure, and antimicrobial activity of ization of δ/γ/σ/γ2/χ phases in silver-doped 2205 duplex stainless steel under
silver(I) complexes constructed with hard and soft donor atoms (n∞{[Ag(L)(PPh3)]2} 800 °C aging, J. Alloys Compd. 633 (2015) 48–53.
with L = α-ala− or asn− and n = 1 or 2), Inorg. Chim. Acta 361 (2008) 1267–1273.
[22] W.-C. Chiang, I.-S. Tseng, P. Møller, L.R. Hilbert, T. Tolker-Nielsen, J.-K. Wu, Influence
of silver additions to type 316 stainless steels on bacterial inhibition, mechanical
properties, and corrosion resistance, Mater. Chem. Phys. 119 (2010) 123–130.

You might also like