current applied physics - buap

9
Effect of Ag doping on the crystallization and phase transition of TiO 2 nanoparticles J. García-Serrano a , E. Gómez-Hernández b , M. Ocampo-Fernández c , U. Pal c, * a Centro de Investigaciones en Materiales y Metalurgia, Universidad Autónoma del Estado de Hidalgo, Carretera Pachuca Tulancingo Km 4.5, Mineral de la Reforma, Hidalgo, C.P. 42184, Mexico b Universidad Popular Autónoma del Estado de Puebla, 21 sur 1103, Col. Santiago, C.P. 72410, Puebla, Pue., Mexico c Instituto de Física, Universidad Autónoma de Puebla, Apdo. Postal J-48, Puebla, Pue. 72570, Mexico article info Article history: Received 21 October 2008 Received in revised form 22 November 2008 Accepted 18 December 2008 Available online 31 December 2008 PACS: 61.46.Hk 61.46.w 64.70.Nd 81.07.b Keywords: TiO 2 nanoparticles Doping Crystallization Phase transition abstract TiO 2 nanoparticles doped with different Ag contents were prepared by a modified sol–gel method, using titanium tetraisopropoxide and silver nitrate as precursors and 2-propanol as solvent. Silver was incor- porated into the TiO 2 matrix via decomposition of AgNO 3 during thermal treatment in different atmo- spheres. Effects of Ag doping on the crystallization and phase transition of the TiO 2 nanoparticles were studied using differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), X-ray diffrac- tion (XRD), and Raman spectroscopy techniques. While air annealing incorporates silver into TiO 2 matrix in silver oxide form, annealing in nitrogen incorporates metallic silver into TiO 2 . Formation of silver oxide increases the thermal stability of the TiO 2 particles. Silver oxide affects the crystallization process of TiO 2 particles and the temperature of transition form anatase to rutile. On the other hand, presence of metallic silver in the samples annealed in nitrogen atmosphere decreases the temperature of phase transition of TiO 2 nanoparticles. Ó 2008 Elsevier B.V. All rights reserved. 1. Introduction In recent years, the preparation and characterization of tita- nium oxide (TiO 2 ) nanostructures have attracted much interest due to their unique properties and potential applications in catal- ysis [1–3], photo-catalysis [4–9], sensors [10,11], solar cells [12–14], energy storage [15], and gene therapy [16,17], among oth- ers. It has been demonstrated that the physical and chemical prop- erties of TiO 2 nanostructures, and hence their potential applications depend strongly on their crystalline structure, mor- phology and particle size [18]. TiO 2 occurs mainly in three crystal- line phases namely anatase, rutile and brookite; and rutile is the thermodynamically most stable phase. They differ in their physical properties, such as refractive index, dielectric constant, chemical and photochemical reactivity. It is well known that the photocata- lytic activity of TiO 2 is intimately related with its crystal structure [19]. While TiO 2 in anatase phase has high photocatalytic activity, the rutile TiO 2 shows no appreciable activity. Also, the surface area, particle size, surface hydroxyl groups, film thickness, and UV light intensity have been seen to be the additional parameters influenc- ing the photocatalytic activity of crystalline TiO 2 [20–22]. On the other hand, the modification of TiO 2 by means of metal doping can also affect the crystallization process, influencing the photocat- alytic efficiency of TiO 2 . Recently, Tomaszewski et al. [23] have re- ported that a reduced sodium concentration in TiO 2 films leads to well crystallized anatase phase after annealing, and increases their photocatalytic activity. Several works have been published on the incorporation of small metal particles, in particular noble metal particles (Ag, Au and Pt) for improving the photocatalytic and photo-electrochemical activities of TiO 2 [24–26]. Addition of small amounts of metal can induce crystallization in amorphous materi- als on annealing at relatively low temperatures. This process, known as metal-induced crystallization, has been applied by Iz- majlowicz et al. [27] for significant reduction of the solid-phase crystallization temperature of amorphous silicon. Crystallization process induced by Au nanoparticles has been studied in thin films of ceramic materials such as SiO 2 , TiO 2 , and ZrO 2 [28]. Very re- cently Perkas et al. [29] have reported that insertion of Au nano- particles into titania by sonochemical reduction of gold ions in ethylene glycol at 80 °C yields anatase phase. There exist many contradictory reports on the crystallization and phase transforma- tion of TiO 2 [30–33]. Though crystallization and phase transforma- tion are essential for many applications involving TiO 2 nanoparticles, it is evident that these can be influenced by the syn- thesis process, starting materials, and impurities or doping species. 1567-1739/$ - see front matter Ó 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.cap.2008.12.008 * Corresponding author. Tel.: +52 222 2295500; fax: +52 222 2295611. E-mail address: [email protected] (U. Pal). Current Applied Physics 9 (2009) 1097–1105 Contents lists available at ScienceDirect Current Applied Physics journal homepage: www.elsevier.com/locate/cap

Upload: others

Post on 27-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Current Applied Physics - BUAP

Current Applied Physics 9 (2009) 1097–1105

Contents lists available at ScienceDirect

Current Applied Physics

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

Effect of Ag doping on the crystallization and phase transition of TiO2 nanoparticles

J. García-Serrano a, E. Gómez-Hernández b, M. Ocampo-Fernández c, U. Pal c,*

a Centro de Investigaciones en Materiales y Metalurgia, Universidad Autónoma del Estado de Hidalgo, Carretera Pachuca Tulancingo Km 4.5,Mineral de la Reforma, Hidalgo, C.P. 42184, Mexicob Universidad Popular Autónoma del Estado de Puebla, 21 sur 1103, Col. Santiago, C.P. 72410, Puebla, Pue., Mexicoc Instituto de Física, Universidad Autónoma de Puebla, Apdo. Postal J-48, Puebla, Pue. 72570, Mexico

a r t i c l e i n f o

Article history:Received 21 October 2008Received in revised form 22 November 2008Accepted 18 December 2008Available online 31 December 2008

PACS:61.46.Hk61.46.�w64.70.Nd81.07.�b

Keywords:TiO2 nanoparticlesDopingCrystallizationPhase transition

1567-1739/$ - see front matter � 2008 Elsevier B.V. Adoi:10.1016/j.cap.2008.12.008

* Corresponding author. Tel.: +52 222 2295500; faxE-mail address: [email protected] (U. Pal).

a b s t r a c t

TiO2 nanoparticles doped with different Ag contents were prepared by a modified sol–gel method, usingtitanium tetraisopropoxide and silver nitrate as precursors and 2-propanol as solvent. Silver was incor-porated into the TiO2 matrix via decomposition of AgNO3 during thermal treatment in different atmo-spheres. Effects of Ag doping on the crystallization and phase transition of the TiO2 nanoparticles werestudied using differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), X-ray diffrac-tion (XRD), and Raman spectroscopy techniques. While air annealing incorporates silver into TiO2 matrixin silver oxide form, annealing in nitrogen incorporates metallic silver into TiO2. Formation of silver oxideincreases the thermal stability of the TiO2 particles. Silver oxide affects the crystallization process of TiO2

particles and the temperature of transition form anatase to rutile. On the other hand, presence of metallicsilver in the samples annealed in nitrogen atmosphere decreases the temperature of phase transition ofTiO2 nanoparticles.

� 2008 Elsevier B.V. All rights reserved.

1. Introduction other hand, the modification of TiO by means of metal doping

In recent years, the preparation and characterization of tita-nium oxide (TiO2) nanostructures have attracted much interestdue to their unique properties and potential applications in catal-ysis [1–3], photo-catalysis [4–9], sensors [10,11], solar cells[12–14], energy storage [15], and gene therapy [16,17], among oth-ers. It has been demonstrated that the physical and chemical prop-erties of TiO2 nanostructures, and hence their potentialapplications depend strongly on their crystalline structure, mor-phology and particle size [18]. TiO2 occurs mainly in three crystal-line phases namely anatase, rutile and brookite; and rutile is thethermodynamically most stable phase. They differ in their physicalproperties, such as refractive index, dielectric constant, chemicaland photochemical reactivity. It is well known that the photocata-lytic activity of TiO2 is intimately related with its crystal structure[19]. While TiO2 in anatase phase has high photocatalytic activity,the rutile TiO2 shows no appreciable activity. Also, the surface area,particle size, surface hydroxyl groups, film thickness, and UV lightintensity have been seen to be the additional parameters influenc-ing the photocatalytic activity of crystalline TiO2 [20–22]. On the

ll rights reserved.

: +52 222 2295611.

2

can also affect the crystallization process, influencing the photocat-alytic efficiency of TiO2. Recently, Tomaszewski et al. [23] have re-ported that a reduced sodium concentration in TiO2 films leads towell crystallized anatase phase after annealing, and increases theirphotocatalytic activity. Several works have been published on theincorporation of small metal particles, in particular noble metalparticles (Ag, Au and Pt) for improving the photocatalytic andphoto-electrochemical activities of TiO2 [24–26]. Addition of smallamounts of metal can induce crystallization in amorphous materi-als on annealing at relatively low temperatures. This process,known as metal-induced crystallization, has been applied by Iz-majlowicz et al. [27] for significant reduction of the solid-phasecrystallization temperature of amorphous silicon. Crystallizationprocess induced by Au nanoparticles has been studied in thin filmsof ceramic materials such as SiO2, TiO2, and ZrO2 [28]. Very re-cently Perkas et al. [29] have reported that insertion of Au nano-particles into titania by sonochemical reduction of gold ions inethylene glycol at 80 �C yields anatase phase. There exist manycontradictory reports on the crystallization and phase transforma-tion of TiO2 [30–33]. Though crystallization and phase transforma-tion are essential for many applications involving TiO2

nanoparticles, it is evident that these can be influenced by the syn-thesis process, starting materials, and impurities or doping species.

Page 2: Current Applied Physics - BUAP

1098 J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105

In the present work, we report on the synthesis of TiO2 and Agdoped TiO2 nanoparticles through a modified sol–gel method,using titanium tetraisopropoxide and silver nitrate as precursorsand 2-propanol as solvent. The silver was incorporated into theTiO2 matrix during the thermal treatment of the gel material inair or N2 via decomposition of AgNO3. The effects of Ag dopingon the morphology, crystallization and phase transition of TiO2

nanoparticles are studied by X-ray diffraction (XRD), scanningelectron microscopy (SEM), Raman and FT-IR spectroscopy, ther-mogravimetric analysis (TGA), and differential scanning calorime-try (DSC) techniques. In general, Ag doping increases the thermalstability of the TiO2 particles when the annealing is performed inair. While Ag doping affects the crystallization process of TiO2 par-ticles, concentration of metallic silver, silver oxide doping and thenature of annealing atmosphere control the temperature of phasetransition from anatase to rutile.

2. Experimental section

2.1. Materials

Titanium tetraisopropoxide (Ti [OCH (CH3)2]4, TTP, 97%) and sil-ver nitrate (AgNO3, 99.99%) were purchased from Aldrich. 2-propa-nol (CH3CHOHCH3, 2P, 99.8%) was purchased from J.T. Baker. Allthe materials were used as received.

2.2. Preparation of TiO2 support

For the preparation of support solution used for the synthesis ofTiO2 and Ag doped TiO2 particles, 14.6 ml of TTP was added drop-wise to 248.8 ml of 2-propanol under vigorous stirring at roomtemperature in a glove box under nitrogen atmosphere. The resul-tant mixture was magnetically stirred for 1 h at room temperatureand then kept in the glove box without agitation for 24 h.

2.3. Preparation of TiO2 nanoparticles

TiO2 nanoparticles were synthesized through a modified sol–gelmethod. First, 131.73 ml of support (TTP–2P) solution was hydro-lyzed by 9.01 ml of water, which was added drop wise under vigor-ous stirring. The resulting sol was stirred for 1 h and maintained inthe glove box for 24 h. Then, the formed gel was dried for 12 h at100 �C in air. The obtained powder sample was heated for 4 h inair or under N2 atmosphere at different temperatures (300, 450,600, 700 and 750 �C), with a heating rate of 240 �C/h.

2.4. Preparation of Ag doped TiO2 particles

The synthesis procedure for Ag–TiO2 nanocomposites was anal-ogous to the one used for TiO2 nanoparticles. However, for Ag dop-ing, different amounts of AgNO3 (0.1698, 0.2831 and 0.4247 g)were added to the reaction mixture. In a typical synthesis process,131.73 ml of the support solution (TTP–2P) was hydrolyzed byslow addition of a solution containing the desired amount ofAgNO3 dissolved in 9.01 ml of water. After that, the resulting mix-ture was maintained in the glove box for 24 h, obtaining a dark yel-low gel. The gels prepared with different AgNO3 contents weredried for 12 h at 100 �C in air, and then treated thermally for 4 hunder air under N2 atmosphere at 300, 450, 600, 700 and 750 �C,with a heating rate of 240 �C/h.

2.5. Characterizations

X-ray diffraction measurements on the samples were per-formed in a Phillips X’Pert diffractometer in the 20–80� range using

Cu Ka (k = 1.5406 Å) radiation. Scanning electron microscopic(SEM) images were recorded with a JEOL JSM LV-5600 microscope.A Nicolet Magna 750 FT-IR spectrometer was used to record the IRabsorption spectra of the samples in the 4000–400 cm�1 spectralrange. For FT-IR studies, 1 mg of solid sample was mixed with99 mg of dry KBr homogeneously to make pellets of 7 mm diame-ter and 0.5 mm thick. A Perkin–Elmer NIR Spectrum GX FT-Ramanspectrometer with Nd:YAG laser source was used for recording Ra-man spectra of the samples at room temperature. Thermogravi-metric measurements were carried out with a Mettler-ToledoTGA/SDTA 851 thermal gravimetric analyzer. The measurementswere performed from room temperature to 1000 �C, at a heatingrate of 10 �C/min under nitrogen atmosphere. Differential scanningcalorimetry measurements were carried out with a Mettler-ToledoDSC-822e calorimeter. The measurements were performed fromroom temperature to 750 �C, at a heating rate of 10 �C/min undernitrogen atmosphere.

3. Results and discussion

Fig. 1 shows the XRD patterns of the undoped TiO2 powdersafter annealing at different temperatures in air for 4 h. The as-pre-pared TiO2 powders were found to be amorphous. The XRD patternof the TiO2 sample annealed at 450 �C revealed 12 peaks in therange of 2h = 25.0 � 76.0�, assigned as (101), (103), (004),(112), (200), (105), (211), (204), (116), (220), (215), and(301) reflections corresponding to the anatase phase. This indi-cates that the amorphous TiO2 nanoparticles crystallized in anatasephase at a temperature below 450 �C. The XRD pattern of the TiO2

sample annealed at 600 �C revealed few rutile peaks in addition tothe peaks corresponding to anatase phase, which indicates that thephase transformation from anatase to rutile starts in the tempera-ture range 450–600 �C. Through semi-quantitative XRD analysis,the fractions of the anatase and rutile phases in this sample wereestimated to be about 47.83% and 52.17%, respectively. For anneal-ing above 600 �C, the relative intensity of the anatase peaks withrespect to rutile peaks decreased and finally vanished. In the XRDpattern of the TiO2 sample annealed at 750 �C, practically onlythe peaks assigned to the rutile phase were observed, indicatinga complete phase transformation from anatase to rutile at thistemperature.

Fig. 2 shows the XRD patterns of the undoped TiO2 powders an-nealed at different temperatures under N2 atmosphere. The as-pre-pared TiO2 powders were found to be amorphous as discussedearlier. The XRD pattern of the TiO2 sample annealed at 450 �C re-vealed 12 peaks in the range of 2h = 25.0 � 76.0�, assigned as(101), (103), (004), (112), (200), (105), (211), (204), (116),(220), (215), and (301) reflections corresponding to the anatasephase, along with 10 peaks assigned as (110), (101), (111),(210), (211), (220), (002), (310), (301), and (112) reflectionsin the same range corresponding to the rutile phase. This showsthat the amorphous TiO2 nanoparticles crystallized in the anataseand rutile phase at a temperature below 450 �C while annealingunder N2 atmosphere. In the XRD pattern of the TiO2 sample an-nealed at 700 �C, the relative intensity of the anatase peaks withrespect to rutile peaks decreased significantly but never disap-peared. The XRD pattern also revealed three more peaks assignedas (002), (301), (202) reflections of the rutile phase, indicatingan incomplete phase transformation from anatase to rutile at thistemperature.

XRD patterns of the Ag–TiO2 composites synthesized with dif-ferent AgNO3 contents before and after annealing in air at differenttemperatures are shown in the Fig. 3. From the diffraction patterns,we can observe: (i) all the as-prepared Ag–TiO2 composites areamorphous; (ii) Ag–TiO2 composites annealed at 450 �C reveal sev-

Page 3: Current Applied Physics - BUAP

20 30 40 50 60 70 80

0

100

200

Inte

nsit

y (a

.u.)

Bragg angle, 2θ (degrees)

as-prepared

0

100

200

300

A (3

01)

A (2

15)

A (2

20)

A (1

16)

A (2

04)

A (2

11)

A (1

05)

A (2

00)

A (0

04)

A (1

12)

A (1

03)A

(101

) 450°C

0

100

200

300

400

500

R (2

02)

A (3

01)

A (2

15)

A (2

20)

R (1

12)

R (3

01)

R (3

10)

R (0

02)

A (2

13)

R (2

20)

A (2

11)

R (2

11)

A (1

05)

R (2

00)

R (2

10)

R (1

11)A

(200

)A

(112

)A

(004

)A (1

03)

R (1

01)R

(110

)

A (1

01) 600°C

0100200300400500600700

R (2

02)

R (1

12)

R (3

01)

R (3

10)

R (0

02)

R (2

20)R

(211

)

R (2

10)

R (1

11)

R (2

00)R

(101

)

R (1

10)

A (1

01)

750°C

Fig. 1. X-ray diffraction patterns of the TiO2 powders before and after annealing at different temperatures in air for 4 h. A and R represents anatase and rutile phases,respectively.

20 30 40 50 60 70 80

020406080

100120140160

R (

310)

R (

112)

0

200

400

600

R (

110)

R (

210)

0

200

400

600

800

1000

A (

200)

R (

202)

R (

311)

R (

301)

R (

002)

R (

211)

A (

211)

A (

101)

A (

103)

A (

004)

A (

112)

A (

200)

A (

105)

A (

211)

A (

204)

A (

220)

A (

215)

A (

301)

as-prepared

450°C

700°C

Inte

nsit

y (a

. u.)

Bragg angle, 2θ (degrees)

R (

101)

R (

111)

A (

200) R (

220)

A (

101) R

(11

0)

R (

101)

R (

200)

R (

111)

R (

210)

R (

211)

R (

220)

R (

002)

R (

310)

R (

301)

R (

112)

R (

215)

A (

004)

Fig. 2. X-ray diffraction patterns of the TiO2 powders before and after annealing at different temperatures in N2 for 4 h. A and R represents anatase and rutile phases,respectively.

J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105 1099

eral diffraction peaks associated to the anatase TiO2 and crystallinesilver oxide compounds, along with some diffraction peaks corre-sponding to metallic silver in face centered cubic phase for high

AgNO3 contents (0.2831 and 0.4247 g of AgNO3). Therefore, theamorphous TiO2 crystallizes in the anatase phase at a temperaturebelow 450 �C and the incorporated Ag in the TiO2 matrix remains

Page 4: Current Applied Physics - BUAP

20 30 40 50 60 70 80

20 30 40 50 60 70 80

20 30 40 50 60 70 80

0

50

100

150

200 as- prepared

Bragg angle, 2θ (degrees)

0

60

120

180

240

Ag2

O3

(571

)

A (

215)

A (2

20)

A (1

16)

A (2

04)

A (

211)

A (1

05)

A (2

00)

Ag2

O3

(440

)

A (1

12)

A (0

04)A (1

01) 450°C

0100200300400500

A (3

01)

A (2

15)

A (2

20)

A (1

16)

R (0

02)

A (2

11)

A (1

05)

A (2

00)

A (1

12)

A (0

04)

A (

103)

R (1

01)

R (1

10)

A (1

01) 600°C

Inte

nsit

y (a

.u.)

0200400600800

1000

R (2

02)

R (

112 )

R (

301)

R (3

10)

R (0

02)

R (2

20)

R (2

11)

R (2

10)

R (1

11)

R (2

00)

Ag

(111

)

Ag2

O3

(131

)

R (1

10)

750°C

0

50

100

150

200 as-prepared

Bragg angle, 2θ (degrees)

0

50

100

150

200 450°C

Ag2

O3

(571

)

A (2

15)

A (2

20)

A (

116)

AgO

(402

)

A (2

04)

Ag

(006

)A

g2O

3 (2

22)

A (1

05)

A (2

00)

Ag2

O3

(440

)

A (0

04)A (1

01)

0

100

200

300

A (2

15)

A (2

20)

R (3

01)

R (0

02)

R (2

20)

Ag2

O (2

20)

R (2

11)

A (1

05)

A (2

00)

R (2

10)

R (1

11)

A (0

04)

R (1

01)

R (1

10)

A (1

01) 600°C

Inte

nsit

y (a

.u.)

0

200

400

600A

g2O

3 (2

62)

R (1

12)

R (3

10)

AgO

(022

)

R (3

10)

R (0

02)

R (2

20)R

(211

)

R (2

10)

R (2

11)

R (2

00)

Ag2

O (2

00)

R (1

01)

R (1

10) 750°C

0

50

100

150

200

Bragg angle, 2θ (degrees)

as- prepared

0

100

200

Ag

(008

)A

(215

)

A (2

20)

AgO

(204

)

A (2

04)

Ag

(006

)A

g2O

3 (1

01)

A (1

05)

A (2

00)

A (0

04)A (1

01)

450°C0

150

300

Ag2

O3

(262

)

A (2

15)

Ag

(200

)A

(116

)

R (3

01)

R (0

02)

Ag2

O3

(711

)A

g (0

06)

R (2

11)

A (2

00)

R (2

01)

R (1

11)

A (0

04)

R (1

01)

R (1

10)

A (

101)

600°C

Inte

nsit

y (a

.u.)

0

200

400

600

800

R (3

01)

AgO

(002

)

AgO

(023

)

Ag2

O3

(660

)

Ag2

O (2

22)

R (3

10)

R (

002)

R (2

20)

Ag2

O3

(260

)

Ag2

O3

(440

)R

(210

)

R (1

11)

R (

200)

Ag2

O (2

00)

R (1

01)

R (

110)

750°C

a

b

c

Fig. 3. X-ray diffraction patterns of the Ag–TiO2 composites synthesized using (a)0.1698 g, (b) 0.2831 g, and (c) 0.4247 g of AgNO3, before and after annealing atdifferent temperatures in air for 4 h. A and R represents anatase and rutile phases,respectively.

1100 J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105

mainly in oxide forms. (iii) On increasing the annealing tempera-ture to 600 �C, the relative intensity of the anatase peaks decreases,and the peaks related to the rutile phase emerges. The intensity ofthe rutile peaks increased with the increase of Ag content in the

composites. Volume fraction of the rutile phase increased from52.17% to 56.25% on incorporating Ag in the TiO2 particles(0.4247 g of AgNO3) during synthesis. This indicates that the pres-ence of Ag species enhances the phase transformation (anatase torutile) process in TiO2. (iv) On annealing at 750 �C, only the peakscorrespond to the rutile and crystalline silver oxide species wereobserved. Absence of any peak characteristic of anatase phase inthe sample suggests that the transformation from anatase to rutilefinishes well below 750 �C while annealing in air atmosphere, dueto the presence of the silver oxide species.

XRD patterns of the Ag–TiO2 composites synthesized with dif-ferent AgNO3 contents with and without annealing treatments atdifferent temperatures in N2 atmosphere are shown in Fig. 4. Fromthe diffraction patterns we can observe: (i) all the as-prepared Ag–TiO2 composites are amorphous; (ii) Ag–TiO2 composites annealedat 300 �C reveals six diffraction peaks associated to the anataseTiO2, while metallic silver is not yet formed. Therefore, the amor-phous TiO2 crystallizes in the anatase phase at a temperature be-low 300 �C in N2 atmosphere. (iii) On increasing the annealingtemperature to 450 �C, the relative intensity of the anatase peaksincreased, the peaks related to the rutile phase did not emerge,and only a small peak corresponding to metallic Ag evolved. (iv)On annealing at 700 �C, the peaks corresponding to anatase, rutileand crystalline Ag species evolved. The low intensity of the rutilepeaks suggests that the transformation from anatase to rutilewas retarded due to the presence silver while annealing in N2

atmosphere.Figs. 5 and 6 show the typical SEM images of TiO2 samples syn-

thesized with and without Ag doping. SEM image (Fig. 5a) of theas-prepared TiO2 shows irregular porous particles of �140 nmaverage size, formed by the aggregation of small spherical particlesof dispersed sizes. Annealing of the TiO2 sample at 750 �C producedmore homogeneous spherical particles, which aggregated to formbigger particles with porous structures (Fig. 5b). Similar morphol-ogy is obtained for the Ag–TiO2 composites annealed at 750 �C(Fig. 6b), whereas the morphology of the as-prepared Ag–TiO2

composites is significantly different (Fig 6a). As-grown Ag–TiO2

composites revealed relatively compact surface morphology withlow porosity in comparison with the undoped as-prepared TiO2

samples (Fig. 5a). In general, the average particle size decreasedon incorporation of Ag in TiO2.

Annealing induced phase transition in TiO2 particles was furthermonitored by Raman spectroscopy. In the Fig. 7, Raman spectra ofthe undoped TiO2 samples annealed at different temperatures inair are presented. While the as-prepared sample showed spectralfeatures basically corresponding to the amorphous phase, the sam-ple annealed at 450 �C revealed several bands at about 206, 396,517, 637, and 669 cm�1. These bands are the characteristic Eg

(low-frequency), B1g, A1g, Eg (high-frequency), and B1g vibrationalmodes of anatase phase [34,35], respectively. With the increaseof annealing temperature to 600 �C, the intensity of these anatasebands increased while two new bands at about 447 and609 cm�1, which correspond to the characteristic Eg and A1g vibra-tional modes of rutile phase, respectively, appeared [36]. For theTiO2 sample annealed at 750 �C, the intensity of the rutile bandsbecame stronger while the anatase bands disappeared, indicatinga complete phase transition from anatase to rutile. Phase transi-tions observed in the Raman spectra are in complete agreementwith our XRD results.

Fig. 8 shows the FT-IR spectra of the TiO2 particles before andafter annealing at 450 �C, along with of as-prepared Ag–TiO2 sam-ples. For the as-prepared TiO2 sample (Fig. 8a) several absorptionbands correspond to the vibrational modes of organic species suchas hydroxyl, carboxilate and alkane groups are observed. A broadband observed in between 3700 and 3000 cm�1 is related to theO–H stretching mode of hydroxyl group, indicating the presence

Page 5: Current Applied Physics - BUAP

0

50

100

Bragg angle, 2θ (degrees)

700º C

450º C

300º C

as- prepared

150

200

250

300

350

A (2

00)

A (2

15)

Ag

(311

)

A (2

15)

Ag

(220

)

A (2

01)

Ag

(200

)

Inte

nsit

y (a

.u.)

Ag

(111

)

A (2

11)

A (2

00)

A (0

04)

A (2

00)

A (2

00)

A (

004)

A (1

05)

A (2

04)

A (1

16)

A (1

05)

A (2

04)

0

50

100

150

200

A (1

01)

A (1

01)

A (1

01)

0

200

400

R (2

00)

R (2

21)

R (1

10)

R (1

01)

R (1

11)

Ag

(200

)

R (

002)

Ag

(220

)

R (

301)

Ag

(311

)

20 30 40 50 60 70 80

20 30 40 50 60 70 80

0255075

100125

R (2

21)

R (1

11)

A (1

03)

A (1

16)

A (2

11)

Inte

nsit

y (a

.u.)

Ag

(311

)

700º C

A (2

20)

R (3

10)

R (0

02)

A (1

05)

R (2

20)A (2

00)

R (2

10)

A (0

04)

R (1

01)

R (1

10)

A (1

01)

A (1

01)

A (0

04)

A (2

00)

A (1

05)

A (2

04)

A (2

20)

450º C

300º C

A (2

20)

A (2

04)

A (1

05)

A (2

00)

A (0

04)

A (1

01)

as- prepared

200

300

400

5000

100

200

A (2

15)

0100200300400500

Bragg angle, 2θ (degrees)

R (3

01)

A (2

11)

A (2

15)

a

b

Fig. 4. X-ray diffraction patterns of the Ag–TiO2 composites synthesized using (a) 0.1698 g, and (b) 0.4247 g of AgNO3, before and after annealing at different temperatures inN2 for 4 h. A and R represents anatase and rutile phases, respectively.

J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105 1101

of moisture in the sample. The peaks in between 3000 and 2800cm�1 are assigned to C–H stretching vibrations of alkane groups.The peak at 1636 cm�1 can be associated to the asymmetricstretching mode of titanium carboxilate. The alkane and carboxi-late groups come from TTP and 2-propanol precursors used inthe synthesis process. The strong absorption band observed in be-tween 800 and 450 cm�1 can bee associated to the vibrationalmodes of TiO2. On annealing at 450 �C the intensity of the bands

associated with the organic groups decreased significantly(Fig. 8b), indicating the elimination of organic species in the sam-ple. On the other hand, all the IR spectra of Ag–TiO2 samples re-vealed a peak at about 1385 cm�1, which was not observed forthe undoped TiO2. The intensity of this peak increased with the in-crease of silver precursor (AgNO3) in the composite samples. Thepeak at 1385 cm�1 was assigned tentatively to the interaction be-tween Ag and TiO2 particles.

Page 6: Current Applied Physics - BUAP

Fig. 5. Typical SEM images of the (a) as-prepared and (b) 750 �C annealed undoped TiO2 samples.

Fig. 6. Typical SEM images of the (a) as- prepared Ag–TiO2 composite synthesized using 0.2831 g of AgNO3, and (b) Ag–TiO2 composite synthesized using 0.1698 g of AgNO3

after annealing at 750 �C.

1102 J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105

The thermal evolution of the TiO2 and Ag–TiO2 samples undernitrogen atmosphere was studied using thermogravimetric analy-sis (TGA) and differential scanning calorimetry (DSC) techniques.Fig. 9 shows typical TGA thermograms of the Ag–TiO2 compositesynthesized using 0.2831 g of AgNO3 and annealed at differenttemperatures in air. Analysis of the TGA thermograms of the as-prepared sample revealed three distinct regions of weight loss.

150 200 250 300 350 400 450 500 550 600 650 700 750 800012340

1

2

A1g

Eg

*

vibrational modes of rutile 750 0C

Inte

nsit

y (a

.u.)

Raman shift (cm-1)

600 0C

0.0

0.5

1.0

1.5 vibrational modes of anataseA

1gB

1g

Eg

Eg

B1g

450 0C

0.0

0.5

1.0 as-prepared

Fig. 7. Raman spectra of the undoped TiO2 particles annealed at differenttemperatures in air.

The first region, a gradual weight loss of about 15% beginning at25 �C and ending at 235 �C, was associated to the loss of residualwater and 2-propanol in the samples. This indicates that althoughthe samples were dried for 12 h at 100 �C in air, water and 2-pro-

4000 3500 3000 2500 2000 1500 1000 500

020406080

100c

Wavenumber (cm-1)

020406080

100 b

Tra

nsm

itta

nce

(%) 0

20406080

100

TiO2 anatase phase

EgCOO- symetricstretch

C-H stretch

a

O-H stretchA2u

Fig. 8. FT-IR spectra of the (a) as-prepared TiO2, (b) TiO2 annealed at 450 �C in air,and (c) as-prepared Ag–TiO2 composites.

Page 7: Current Applied Physics - BUAP

0 100 200 300 400 500 600 700-10

-5

0

5

10

15

290 0C

598 0C

437.6 0C

190 0C

98 0C

Hea

t fl

ow (

mW

)

Temperature (° C)

Fig. 10. DSC curve of as-prepared TiO2 particles. The measurement was performedfrom room temperature to 750 �C, at a heating rate of 10 �C/min under nitrogenatmosphere.

200 400 600 800 1000 75

80

85

90

95

100 dc

b

a

1000 ºC; 78.63%

+

+

+

402.7 ºC; 79.98%

235.4 ºC, 85.07%

Wei

ght

(%)

Temperature (°C)

Fig. 9. TGA thermograms of the Ag–TiO2 composites synthesized using 0.2831 g ofAgNO3; (a) as-prepared, annealed at: (b) 450 �C, (c) 600 �C, and (d) 750 �C in air.

J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105 1103

panol were not eliminated completely until about 235 �C temper-ature. In the second region, about 5% of weight was lost between230 and 403 �C. Such weight loss can be associated to the elimina-tion of organic components. Finally, the last weight loss occurredabove 403 �C, which represents approximately 1% of initial solidweight.

On the other hand, as expected for the Ag–TiO2 composite sam-ples annealed at 450 �C or higher, the weight losses associated tothe elimination of water and organic components decreased or ab-sent in their TGA thermograms. The TGA curves of the Ag–TiO2

composites annealed at 450, 600, and 750 �C revealed only oneweight loss throughout the whole measurement range, represent-ing approximately 3%, 1.2%, and 0.7% of the initial weight, respec-tively, indicating the elimination of practically all water andorganic species by the thermal treatments before realizing theTGA analysis. The results are in well agreement with the IR analysispresented earlier. The TGA thermograms of Ag–TiO2 compositessynthesized with 0.1698 g and 0.4247 g of AgNO3 revealed similarcharacteristics as that of Ag–TiO2 composite prepared with0.2831 g of AgNO3.

DSC analysis was performed to determine the effect of silverdoping on the crystallization and phase transformation behaviorof the TiO2 nanoparticles. Fig. 10 shows the DSC curve of the as-prepared TiO2 particles. The DSC curve revealed an endothermicpeak centered at about 98 �C and four exothermic peaks locatedat about 190, 290, 437, and 598 �C. The endothermic peak wasattributed to the elimination of water adsorbed on the surface ofthe TiO2 particles. The broad nature of the endothermic peak leadsto the conclusion that the dehydration in TiO2 is a slow process. Onthe other hand, XRD analysis has shown that the TiO2 particlesheated in air at 100 �C were amorphous in nature and after anneal-ing at 450 �C they transform to anatase phase completely. There-fore, the first exothermic peak observed at about 190 �C can beattributed to the crystallization of amorphous TiO2 to anatasephase, indicating that the TiO2 nanoparticles obtained by our mod-ified sol–gel method crystallize at relatively lower temperature.The second exothermic peak at about 290 �C was not well resolvedin the DSC measurement. However, this temperature correspondsto the temperature associated to the elimination of organic compo-nents as observed in the TGA thermogram (Fig. 9). From TGA ther-mogram it can also be observed that the process associated withthe organic species elimination is a complicated one. These resultssuggest that the elimination of organic compounds does not occur

by means of a simple desorption process, but also involves thechemical decomposition of the species during a combustion pro-cess. Therefore, the DSC exothermic peak at 290 �C was associatedto the combustion process of the organic species. On the otherhand, the exothermic peak observed at about 437 �C was attributedto the phase transformation from anatase to rutile, in agreementwith the previously reported results [37]. We assigned the exother-mic peak at about 598 �C to the rutile phase formation. It should beremembered that the XRD and Raman results (Figs. 1 and 7) of theundoped TiO2 particles revealed the occurrence of anatase–rutilephase transition in between 450 and 750 �C. It is important to notethat while our XRD and Raman analysis revealed only the anatasephase after annealing the sample at 450 �C, the DSC curve of thesample indicates that the phase transformation starts even below450 �C. To verify this point, we performed a DSC analysis of thesample after annealing at 450 �C. However, no peak correspondingto anatase–rutile transition was observed for the sample. This sug-gests in fact the anatase–rutile transition starts at temperature be-low 450 �C, which could not be detected from the XRD and Ramanspectra of the sample.

As the nucleation and growth in solid state can occur throughdifferent rate controlled processes like diffusion and/or phaseboundary reaction, a plausible explanation for the discrepanciesbetween the XRD and DCS results can be given by consideringthe phase transformation process in our nanostructures. Probably,the transformation from anatase to rutile phase in our samples oc-curs through nucleation and growth of rutile phase in the bulk, e.g.from the center of the particles, beginning at about 437 �C. Then,the growth of rutile propagates slowly from the centre of the par-ticles to the surface, as shown in the Fig. 11. The formation of rutilephase can be detected by XRD when it is extended near to the sur-face of the particles. Therefore, until a good fraction of the particlesdoes not transform to rutile, it could not be detected in their XRDpatterns.

In the Fig. 12, the DSC curves of the Ag–TiO2 composite particlessynthesized using three different amounts of AgNO3 are presented.From the DSC curves, we can observe an endothermic peak atabout 98 �C for all the three samples, which correspond to thedehydration process. The peak related to organic species elimina-tion can be found at about 290 �C for all the Ag–TiO2 samples.The fact that the peaks attributed to the dehydration process and

Page 8: Current Applied Physics - BUAP

0 300 600-0.5

0.0

0.5

1.0

472.9 0C

454.7 0C

without Ag 0.1698 Ag 0.4247 Ag

423.2 0C

Temperature (ºC)

Hea

t fl

ow (

mW

)

171 0C

166 0C

141 0C

Fig. 13. DSC curves Ag–TiO2 composites prepared without silver, with 0.1698 g and0.4247 g of AgNO3 and annealed at 300 �C in nitrogen atmosphere for 4 h. Themeasurements were performed from room temperature to 700 �C, at a heating rateof 10 �C/min under nitrogen atmosphere.

0 100 200 300 400 500 600 700-10

-5

0

5

10

15

461 0C

491 0C216.8 0C

Hea

t flo

w (

mW

)

Temperature ( °C)

a

b

c

d

Fig. 12. DSC curves of as-prepared sol–gel grown Ag–TiO2 composites preparedwith (a) 0.1698 g, (b) 0.2831 g, and (c) 4247 g of AgNO3. The measurements wereperformed from room temperature to 750 �C, at a heating rate of 10 �C/min undernitrogen atmosphere.

Fig. 11. Schematic presentation of the proposed phase transformation process in the sol–gel grown TiO2 nanoparticles: (a) particles of TiO2 crystallized with anatase phase;(b) nucleation of rutile crystals at around 437 �C; (c) growth of rutile crystals at temperatures in between 437 and 550 �C; (d) growth of rutile crystals close to the surface ofthe particles at 600 �C; (e) complete transformation of the crystals to the rutile phase at 750 �C.

1104 J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105

decomposition of organic species were observed at the same tem-peratures (98 and 290 �C, respectively) in both TiO2 and Ag–TiO2

nanoparticles indicates that these processes are independent ofAg doping. The peak associated with the crystallization of amor-phous TiO2 to anatase phase was observed at about 217 �C. Theshift of the peak towards higher temperatures with respect to itsposition in undoped TiO2 could be associated to the Ag doping. Incontrast to the undoped TiO2 sample, where the DSC peak corre-sponds anatase–rutile transformation appeared at 437 �C, theDSC curve of the Ag–TiO2 sample prepared with 0.1698 g of AgNO3

revealed this peak at about 461 �C. The peak associated with theanatase–rutile transformation shifted towards higher temperaturewith the increase of the used AgNO3 amount in the synthesis pro-cess. The peak shifted up to 469 �C and 491 �C for the samples syn-thesized using 0.2831 g and 0.4247 g of AgNO3, respectively. It isevident that the thermal stability of the anatase phase increaseswith the increase Ag doping level of the TiO2 particles. It must benoted that our observations are in contradiction to the previouslyreported results of Chao et al. [38] where the anatase–rutile trans-formation for silver doped titania was reported to occur at lowertemperatures.

In the Fig. 13, the DSC curves of the Ag–TiO2 composite particlesannealed at 300 �C in nitrogen atmosphere for 4 h are shown. Fromthe DSC curves, we can observe an endothermic peak at about98 �C for all the three samples, which corresponds to the dehydra-tion process. The peak associated with the crystallization processof titania was observed at about 171 �C for the sample preparedwithout Ag, at about 166 �C for the sample prepared with0.1698 g of AgNO3 and at about 141 �C for the sample preparedwith 0.4247 g of AgNO3; which suggests that a thermal treatmentat 300 �C in nitrogen atmosphere for 4 h was enough to complete

the crystallization process. In addiction, the shift of the peak to-wards lower temperatures with respect to its position for undopedTiO2 could be associated to the Ag doping and nitrogen atmo-sphere. Also, we can see a peak of anatase–rutile transformationat about 423 �C in the DSC curve of the undoped TiO2 sample an-nealed at 300 �C in nitrogen atmosphere, in contrast to the DSCcurve of the undoped TiO2 sample where it appeared at about437 �C. Similarly, the peak associated with the anatase–rutiletransformation shifted towards lowers temperatures for the TiO2

samples prepared with 0.1698 g and 0.4247 g of AgNO3 and an-nealed at 300 �C under nitrogen in comparison with the un-an-nealed samples and the samples annealed in air. Therefore, thethermal stability of the anatase phase decreases when the samplesare treated in nitrogen atmosphere.

4. Conclusions

TiO2 nanoparticles doped with different Ag contents were syn-thesized through a modified sol–gel technique using titanium tet-raisopropoxide and silver nitrate as precursors in a 2-propanolsolution. Incorporation of silver in the TiO2 particles occurs duringthe thermal treatment of the composite gel material throughdecomposition of AgNO3. The XRD results indicate that when thesamples are treated in air, the silver incorporates as silver oxide,while on thermal treatment in nitrogen ambient it incorporatesas metallic silver. The Ag doping increases the thermal stabilityof the TiO2 particles when annealed in air, enhancing the crystalli-zation process, and inhibiting the normal anatase–rutile phasetransition. It is predicted that the anatase–rutile phase transitionstarts from the centre of the nanoparticles, which extends to the

Page 9: Current Applied Physics - BUAP

J. García-Serrano et al. / Current Applied Physics 9 (2009) 1097–1105 1105

particle surface on prolonged thermal treatment at high tempera-tures. On the other hand, for the Ag–TiO2 composite particles an-nealed in N2 atmosphere, the transformation from anatase torutile starts at temperatures below 450 �C, due to the presencemetallic silver nanoparticles. Even on annealing above 600 �C, acomplete anatase–rutile phase transformation does not occur asin the case of the samples annealed in air. Therefore, the thermalstability of the anatase phase depends strongly both on the Ag dop-ing and the annealing ambient.

Acknowledgements

The work was partially supported by CONACyT, Mexico (Grant# 46269) and VIEP-BUAP (Grant # 93/EXC/2008-1), Puebla.

References

[1] K.I. Hadjiivanov, D.G. Klissurski, Chem. Soc. Rev. 25 (1996) 61.[2] P. Forzatti, Catal. Today 62 (2000) 51.[3] Z. Ma, Y. Yue, X. Deng, Z. Gao, J. Mol. Catal. A 178 (2002) 97.[4] C.A. Linkous, G.J. Carter, D.B. Locuson, A.J. Ouellette, D.K. Slattery, L.A. Smitha,

Environ. Sci. Technol. 34 (2000) 4754.[5] J.K. Yang, A.P. Davis, Environ. Sci. Technol. 34 (2000) 3796.[6] K.J. Buechler, R.D. Noble, C.A. Koval, W.A. Jacoby, Ind. Eng. Chem. Res. 38 (1999)

892.[7] A. Maldoti, A. Molinari, R. Amadeni, Chem. Rev. 102 (2002) 3811.[8] M. Anpo, M. Takeuchi, J. Catal. 216 (2003) 505.[9] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. C: Photochem. Rev. 1

(2000) 1.[10] I. Hayakawa, Y. Iwamoto, K. Kikuta, S. Hirano, Sens. Actuators, B: Chem. 62

(2000) 55.[11] Y. Zhu, Anal. Chem. 74 (2002) 120.[12] U. Bach, D. Lupo, P. Comte, J.A. Moser, F. Weissortel, J. Salbeck, H. Spreitzer, M.

Gratzel, Nature 395 (1998) 583.[13] M. Gratzel, Nature 414 (2001) 338.

[14] M. Adachi, J. Am. Chem. Soc. 126 (2004) 14943.[15] S. Zhang, L.M. Peng, Q. Chen, G.H. Du, G. Dawson, W.Z. Zhou, Phys. Rev. Lett. 91

(2003) 256103.[16] K. Mier, M. Gratzel, Chem. Phys. Chem. 3 (2002) 371.[17] T. Paunesku, T. Rajh, G. Wiederrecht, J. Master, S. Vogt, N. Stojicevic, M. Protic,

B. Lai, J. Oryhon, M. Thurnauer, G. Woloschak, Nat. Mater. 2 (2003) 343.[18] W. Wang, B. Gu, L. Liang, W.A. Hamilton, D.J. Wesolowski, J. Phys. Chem. B 108

(2004) 39.[19] B. Ohtani, Y. Ogawa, S. Nishimoto, J. Phys. Chem. B 101 (1997) 3746.[20] B. Ohtani, S. Zhang, S. Nishimoto, T. Kagiya, J. Photochem. Photobiol. A: Chem.

64 (1992) 223.[21] K. Tanaka, M.F.V. Capule, T. Hisanaga, Chem. Phys. Lett. 187 (1991) 73.[22] M. Fernández-Garcia, A. Martinez Arias, J.C. Hanson, J.A. Rodriguez, Chem. Rev.

104 (2004) 4063.[23] H. Tomaszewski, K. Eufinger, H. Poelmen, D. Poelman, R. Gryse, P.F. Smet, G.B.

Marin, Int. J. Photoenergy 2007 (2007) 5p. (Article ID95213).[24] F. Boccuzzi, A. Chiorino, S. Tsubota, M. Haruta, J. Phys. Chem. 100 (1996) 3625.[25] Z. Zhang, C.C. Wang, R. Zakaria, J.Y. Ying, J. Phys. Chem. B 102 (1998) 10871.[26] H. Kominami, S. Muratami, J. Kato, Y. Kera, B. Ohtani, J. Phys. Chem. B 106

(2002) 10501.[27] M.A.T. Izmajlowicz, A.J. Flewitt, W.I. Milne, N.A. Morrison, J. Appl. Phys. 94

(2003) 7535.[28] M. Epifani, C. Giannini, L. Tapfer, L. Vasanelli, J. Am. Ceram. Soc. 83 (2000)

2385.[29] N. Perkas, V.G. Pol, S.V. Pol, A. Gedanken, Cryst. Growth Des. 6 (2006) 293.[30] B.D. Stojanovic, Z.V. Marinkovic, G.O. Brankovic, E. Fidancevska, J. Therm. Anal.

Calorim. 60 (2000) 595.[31] P.I. Gouma, M.J. Mills, J. Am. Ceram. Soc. 84 (2001) 619.[32] O.K. Varghese, D. Gong, M. Paulose, C.A. Grimes, E.C. Dickey, J. Mater. Res. 18

(2003) 156.[33] M. Fernandez-Garcia, X. Wang, C. Belver, J.C. Hanson, J.A. Rodriguez, J. Phys.

Chem. C 111 (2007) 674.[34] A.G. Tompsett, A.G. Bowmaker, P.R. Gooney, B.J. Metson, A.K. Rodgers, M.J.

Seakins, J. Raman Spectrosc. 26 (1995) 57.[35] T. Ohsaka, F. Izumi, Y. Fujiki, J. Raman Spectrosc. 7 (1978) 321.[36] S.P.S. Porto, P.A. Fleury, T.C. Damen, Phys. Rev. 154 (1967) 522.[37] C. Xie, Z. Xu, Q.J. Yang, BY. Xue, Y.G. D, J.H. Zhang, Mater. Sci. Eng. B 112 (2004)

34.[38] H.E. Chao, Y.U. Yun, H.U. Xiangfang, A. Larbot, J. Eur. Ceram. Soc. 23 (2003)

1457.