CHAPTER 4 SYNTHESIS, STRUCTURAL AND OPTICAL...

34
67 CHAPTER 4 SYNTHESIS, STRUCTURAL AND OPTICAL PROPERTIES OF SOME RARE EARTH (Eu, Ce, Sm) DOPED CADMIUM SULFIDE NANOPARTICLES 4.1 INTRODUCTION Semiconductor nanostructures have drawn much attention due to their unique mechanical, optical, and electronic properties. Inorganic compounds doped with lanthanide ions (Stouwdam et al 2003) are widely used as the luminescent materials in lighting and displays (Blasse and Grabmaier 1994) optical amplifiers (Digonet 2001) and lasers (Reisfeld and Jorgensen 1977). Recently, optical properties of impurity doped nanocrystals have attracted much attention as they are expected to modify both the electronic states and electromagnetic fields. Therefore, a possible influence of quantum size effect on the luminescence properties is expected for II–VI semiconductor nanocrystals due to the inclusion of rare earth (RE) metal ions (Konishi et al 2001). The confinement quantum size effects of semiconductor nanoparticles not only create photogenerated carriers, which may have an interaction with f-electrons but also significantly influences the optical properties (Chowdhury et al 2004). Excitonic emission in the host and an improvement in the luminescence intensity are also expected for the semiconductor nanostructures after the RE doping. The effects of the reduced dimensionality on the electronic relaxation and the phonon density-of-states of semiconductor and insulating nanocrystals has been extensively

Transcript of CHAPTER 4 SYNTHESIS, STRUCTURAL AND OPTICAL...

  • 67

    CHAPTER 4

    SYNTHESIS, STRUCTURAL AND OPTICAL PROPERTIES

    OF SOME RARE EARTH (Eu, Ce, Sm) DOPED CADMIUM

    SULFIDE NANOPARTICLES

    4.1 INTRODUCTION

    Semiconductor nanostructures have drawn much attention due to

    their unique mechanical, optical, and electronic properties. Inorganic

    compounds doped with lanthanide ions (Stouwdam et al 2003) are widely

    used as the luminescent materials in lighting and displays (Blasse and

    Grabmaier 1994) optical amplifiers (Digonet 2001) and lasers (Reisfeld and

    Jorgensen 1977). Recently, optical properties of impurity doped nanocrystals

    have attracted much attention as they are expected to modify both the

    electronic states and electromagnetic fields. Therefore, a possible influence of

    quantum size effect on the luminescence properties is expected for II–VI

    semiconductor nanocrystals due to the inclusion of rare earth (RE) metal ions

    (Konishi et al 2001). The confinement quantum size effects of semiconductor

    nanoparticles not only create photogenerated carriers, which may have an

    interaction with f-electrons but also significantly influences the optical

    properties (Chowdhury et al 2004). Excitonic emission in the host and an

    improvement in the luminescence intensity are also expected for the

    semiconductor nanostructures after the RE doping. The effects of the reduced

    dimensionality on the electronic relaxation and the phonon density-of-states

    of semiconductor and insulating nanocrystals has been extensively

  • 68

    investigated by both theoretical and experimental approaches (Yang et al

    2000, Simon and Geller 2001).

    The spontaneous emission probability of optical transitions

    (luminescence lifetime) from RE ions doped semiconductor nanostructures

    may be significantly different from the bulk counterparts. In order to improve

    the luminescence properties of CdS nanoparticles, lanthanide ions like Eu3+

    could be incorporated. Europium doped fluorescent nanocrystals have been

    extensively studied because of their properties for application in lighting and

    display phosphors. It is possible that Eu3+ can substitute Cd2+ because of

    similar ionic radii (ionic radii of Eu3+ and Cd2+ are 0.94 Å and 0.97 Å

    respectively). It is expected that 4f electrons in the rare earth metal ions

    participate in luminescence. These are hardly influenced by their ligands due

    to the presence of 5s and 5p electrons surrounding them. Therefore, crystal

    field effects observable in 3d transition metal ions are not feasible in RE

    metal ions. However, rare earth ion doped phosphors have emission in the

    visible range.

    As the optical and electronic properties of the semiconductor

    nanocrystals are significantly influenced by the RE doping, recently much

    attention has been paid for the synthesis of RE doped semiconductor

    nanostructures. Yang et al (2004) synthesised europium doped ZnS

    nanocrystals and reported that the doped materials exhibit much better

    luminescent properties than the pure ZnS nanocrystals. Ageeth et al (2002)

    fabricated the europium doped CdS nanocrystals by both the micro-emulsion

    and the precipitation methods. Zhu et al (2006) demonstrated that Eu3+ on the

    surface of the semiconductor nanoparticle significantly increases the

    fluorescence intensity of band gap emission.

    Doping of II–VI compounds with trivalent ions like cerium,

    terbium and europium etc., has been extensively studied (Okamoto et al 1988,

  • 69

    Jayaraj and Vallabhan 1991). Doping of rare earth element like cerium

    reduces the particle size of nanomaterials and also increases surface area

    (Liqiang et al 2004 and Guo et al 2005). Recently, Vij et al (2009) and

    Sharma et al (2009) reported the luminescence studies on Ce doped SrS

    nanostructures and Ce doped CaS nanoparticles synthesized by solid state

    diffusion method. Vinay Kumar et al (2010) reported luminescence

    investigations on Ce3+ doped CaS synthesized using the chemical

    co-precipitation method.

    Spectroscopic studies of trivalent samarium ions have received

    much interest because of good fluorescence efficiency in the visible and

    infrared region (Lin and Pun 2002). The understanding of the optical

    properties of Sm3+ ions is of great importance due to its potential

    technological applications (Nemec et al 2003). Mathew et al discussed the

    optical (Mathew et al 2008) and dielectric properties (Mathew et al 2008a)

    CdTe/Sm3+ in sol–gel silica glasses synthesised by sol-gel method. ZnS:Sm is

    known to be a good red emitter in thin film electroluminescence devices

    (Tohda et al 1986), and optical properties of Sm-doped ZnS bulk crystals

    have been studied by several investigators (Swiatek et al 1991 and Hommel et

    al 1991). Xu et al (2002) also reported that doping of La3+, Ce3+, Er3+, Pr3+,

    Gd3+, Nd3+ or Sm3+ with TiO2 was beneficial for NO2 adsorption to enhance

    the photocatalytic activity.

    Wet chemical processes are practical approaches for the synthesis

    of metal doped nanoparticles in a bulk quantity. Although the doping process

    via wet chemical synthesis offers excellent electronic and magnetic properties

    in the semiconducting nanostructures, the perfect size, structure, and shape

    control of the doped semiconductor nanostructures remain a big challenge as

    they significantly influence their electronic, magnetic, and optical properties.

  • 70

    50 ml of 0.652 g cadmium acetate in mixed solvent 50 ml of 0.008 g europium acetate in mixed solvent

    Stirring at room temperature for 2 h

    Dropwise addition of 50 ml solution of 0.195 g sodium sulphide

    Ultrasonication for 60 minutes

    washed with DI water and ethanol

    Separation of nanoparticles by centrifugation at 4000 rpm

    Dried in a hot air oven for 24 h

    50 ml of 0.652 g cadmium acetate in mixed solvent 50 ml of 0.008 g europium acetate in mixed solvent

    Stirring at room temperature for 2 h

    Dropwise addition of 50 ml solution of 0.195 g sodium sulphide

    Ultrasonication for 60 minutes

    washed with DI water and ethanol

    Separation of nanoparticles by centrifugation at 4000 rpm

    Dried in a hot air oven for 24 h

    Herein we report the preparation of europium doped CdS

    nanostructures by co-precipitation method with isopropyl alcohol (IPA) and

    ethylene glycol (EG) as solvents. We also demonstrate that the size, structure,

    morphology and the optical properties of the europium doped CdS

    nanostructure can be finely controlled by simply varying the nature of the

    solvent used in the synthesis. We also investigate the effect of cerium and

    samarium ion doping on the structural and optical properties of CdS

    nanoparticles synthesized by chemical co-precipitation method.

    4.2 Eu3+

    DOPED CdS NANOPARTICLES

    4.2.1 Materials Synthesis

    The starting materials for the synthesis of Eu3+ doped cadmium sulfide

    nanocrystals (Cd1−xEuxS with x = 2%) were (CH3COO)2Cd · 2H2O, Eu(CH3COO)3 ·

    H2O, Na2S, which are of analytic grade. Isopropyl alcohol (IPA) and ethylene glycol

    (EG) were purchased from CDH, India. Solvents were prepared by mixing equal

    volume of DI water with the respective solvents such as IPA and EG. The solution

    (50 ml) of 0.652 g cadmium acetate and 50 ml solution of 0.008 g europium acetate

    were prepared separately using EG-DI water as mixed solvent. The solution of

    europium acetate was added slowly into the cadmium acetate solution with a

    constant stirring. The mixture was subjected to stirring at room temperature for 2 h

  • 71

    in order to achieve a good dispersion of Cd2+ and Eu3+ into the solvent. With this

    mixture 50 ml solution of 0.05 M sodium sulfide (0.195 g) in EG-DI water was

    added drop wise and it was observed that the mixed solution turned to pale yellow

    initially and changes into yellowish orange color after stirring for an hour. The

    precipitate obtained was subjected to ultrasonication for 60 min. and washed with

    ethanol and DI water for four times in order to remove insoluble residues and

    impurities. The nanoparticles were separated from solution by centrifugation (4000

    rpm) and dried in a hot air oven at 60 C. The above procedure shown in flow chart was followed for the second reaction except that mixture of IPA-DI water was used

    as a mixed solvent instead of EG. The sample prepared by using EG-DI mixed

    solvent was denoted as Eu:CdS-1, whereas the sample prepared by using IPA-DI

    was named as Eu:CdS-2.

    4.2.2 Characterisation Methods

    X-ray diffraction patterns were recorded using PANalytical X-ray

    diffractometer with CuK radiation ( = 1.5406 Å) in the range of 20 to 60 (2)at a scanning rate of 0.05 /min. Morphology and composition of the synthesised samples was observed with a Hitachi S-3400 SEM-EDAX and confirmed with

    Nippon Jarrell-Ash, IRIS Advantage ICP-OES spectrophotometer. High resolution

    transmission electron microscope (HRTEM) and SAED studies were performed at

    200 keV using JEOL JEM 3010 with LaB6 filament. UV-reflectance studies were

    carried out using CARY 5E UV-VIS Reflectance mode-UV spectrophotometer.

    XPS measurements were carried out using PHI Quantera SXM (ULVAC-PHI)

    with Monochromatic Al Kα as an X-ray source. The specific surface area was

    measured on a Quantachrome Autosorb-1 adsorption analyzer using Brunauer-

    Emmett-Teller (BET) method. Photoluminescence spectra were recorded using

    Shimadzu-5301 spectroflurometer with 450W Xenon lamp source. Raman

    measurement were analysed with Horiba Jobin-Yvon T64000 photon design micro

    Raman spectrophotometer using Ar ion laser with the excitation of 514.5 nm.

  • 72

    4.2.3 Results and Discussion

    4.2.3.1 X-ray Diffraction Analysis

    Figure 4.1 shows the powder XRD patterns of europium doped

    cadmium sulfide nanocrystals prepared by using EG-DI water and IPA-DI water

    as mixed solvents. It is observed from the XRD patterns that the samples are

    highly crystalline but with different structure. The materials prepared with EG-DI

    water exhibits several peaks which can be indexed to (100), (002), (101) and (102)

    planes of the hexagonal wurtzite CdS, whereas for Eu:CdS-2 prepared using IPA-

    DI water shows (111), (220) and (311) peaks indexed to cubic (zinc-blende)

    structure. In addition, the crystallite size of the sample is significantly changed by

    changing the mixed solvents. From the Scherrer equation the average crystallite

    sizes for Eu:CdS-1 and Eu:CdS-2 were calculated to be 3.65 and 2.64 nm

    respectively. This indicates that IPA-DI water solvent plays a significant role in

    reducing the crystallite size, which is mainly due to the fact that density and the

    viscosity of the solvent is much lower than that of EG-DI water. From these

    results, it can be concluded that it is extremely important to have a low viscous

    solvent with less density for the preparation of CdS crystals with small size.

    It is also observed that the doping of CdS with europium does not

    make much change in the structure of the CdS because no shift in the peak

    positions was observed for both the sample irrespective of the usage of the

    solvent. This is may be due to the doping of small percentage of europium in

    the CdS lattice. However, the width of the XRD peaks is broadened after the

    europium doping, revealing the decrease in the crystallite size. The absence of

    peaks other than CdS confirms the purity of the sample. These results further

    indicate that the dopant ions are incorporated into lattice of nanocrystals.

    Similar results on the successful incorporation of the Eu3+ ion in the CdSe

    nanocrystals with random ion displacement of the cadmium cation site were

    also reported (Raola and Strouse 2002).

  • 73

    Figure 4.1 Powder X-ray diffraction of as synthesised europium doped

    CdS nanostructures

    4.2.3.2 Structural, Morphological and Compositional Analysis

    Figure 4.2 (a)–(d) show the HRSEM images with different

    magnifications of Eu:CdS-1 and Eu:CdS-2 samples respectively. A significant

    change in the morphology was observed for the samples prepared using

    different mixed solvents. Flower like morphology was observed for Eu:CdS-1

    whereas clusters of rice like morphology was obtained for Eu:CdS-2. The

    HRSEM images also show that the particles are highly uniform and

    distributed homogeneously without any agglomeration for Eu:CdS-2. In

    addition, particles are clear and each rice like particle is connected and

    appears as a group of particle for the same samples. However, several

    agglomerated particles with different size were observed for Eu:CdS-1, which

    is mainly due to the high viscous nature of the EG. From these results, it is

    concluded that the size, shape and agglomeration of the crystalline particles

    can be controlled by simply varying the mixed solvent system. The doping of

    CdS nanostructures with europium was also confirmed by ICP analysis and

    the amount of europium present in the final product is given in Table 4.1.

  • 74

    Figure 4.2 HRSEM images of europium doped cadmium sulfide

    nanostructures (a) and (b) Eu:CdS-1 and (c) and (d)

    Eu:CdS-2

    Table 4.1 Elemental Composition of Eu:CdS Nanostructures Calculated

    from ICP Analysis

    ElementObserved value (mg/l) Calculated percentage

    Eu:CdS-1 Eu:CdS-2 Eu:CdS-1 Eu:CdS-2

    Cd 361.60 575.9 83.2933 65.55115

    S 19.04 82.9 15.3776 33.08485

    Eu 7.80 16.2 1.3290 1.36399

  • 75

    4.2.3.3 High Resolution Transmission Electron Microscopy

    Figure 4.3 (a) and (c) show the HRTEM images of Eu:CdS

    nanocrystals of different structures indicating the shape of the particles is

    approximately spherical, with slight prolate deviations similar to the earlier

    report (Tong and Zhu 2006). The particles are connected each other and create

    the nanospace between the particles which may give them excellent textural

    properties. The mean diameter was about 5 nm determined from the HRTEM

    image, which is slightly larger than the value estimated from the XRD

    analysis. The crystalline domain size calculated from XRD patterns are much

    smaller than those obtained from HRTEM, because of the existence of crystal

    defect and lattice distortion resulted from doping (Qu et al 2002). Figure 4.3

    (b) and (d) show the lattice fringe pattern of the synthesized nanocrystals,

    confirming the highly crystalline nature of the final product. HRTEM images

    of Cd0.98Eu0.02S illustrate the atomic resolution of these particles (showed in

    round circle) signifying the particles are in nanodimension. Insets in Figure

    4.3 (b) and (d) show the SAED pattern of Eu:CdS-1 and Eu:CdS-2

    nanostructures respectively. The electron diffraction pattern on a number of

    nanocrystals, consists of broad diffuse rings further confirming the crystalline

    nature of the Eu:CdS nanostructures.

  • 76

    Figure 4.3 (a) and (b) HRTEM images of the sample Eu:CdS-1

    nanostructures and (c) and (d) HRTEM images of the sample

    Eu:CdS-2 nanostructures.

    4.2.3.4 UV-reflectance Studies

    UV-reflectance spectra for the synthesised europium doped CdS

    nanostructures are shown in Figure 4.4. The band gap values of the Eu:CdS-1

    and Eu:CdS-2 were estimated with the well defined excitonic peaks at 444 nm

    (2.79 eV) and 462 nm (2.68 eV) respectively by extrapolating the reflectance

    spectra. It is known that the excitonic absorption peak is associated with the

    lowest optical transition and provides a simple way to determine the bandgap

  • 77

    of nanocrystals. A clear blue-shift is observed in the reflectance spectra. This

    shift towards the shorter wavelength indicates the increase of optical band

    gap. The observed difference in the band gap value for both the samples

    may due to the variation in the crystallite size. It is reported for CdS that

    quantum effects are more pronounced in the size range from 2 nm to 10 nm

    (Brus 1986). The radius (R) of CdS NPs can be calculated using the effective

    mass approximation model given in the Equation 4.1.

    R

    2e8.1

    hm

    1

    em

    12R2

    22hE

    (4.1)

    where ΔE = 0.37 eV (Eu:CdS-1) and 0.26 eV (Eu:CdS-2) is the increase of the band gap energy, ε = 5.7 (Li and Du 2003) is the relative dielectric constant, and me = 0.19mo and mh = 0.8mo are the effective masses of

    electrons and holes respectively, where mo is the free electron mass. The

    calculated particle size was 3.59 nm for Eu:CdS-1 and 2.43 nm for Eu:CdS-2

    which are in good agreement with the crystallite size calculated from the

    XRD data.

    Figure 4.4 Reflectance mode UV spectra of Eu doped CdS

    nanostructures.

  • 78

    Figure 4.5 Photoluminescence spectra of Eu doped CdS nanostructures

    (inset shows photoluminescence spectra for pure CdS).

    4.2.3.5 Photoluminescence Studies

    Figure 4.5 shows the PL spectra of the synthesised Eu3+ doped CdS

    nanostructures excited at 350 nm. Inset figure shows the emission spectrum of

    pure cadmium sulfide nanocrystals. Compared to the pure CdS, intense emission

    peaks at 475 and 574 nm was observed for Eu3+ doped CdS. The emission peak at

    574 nm, due to the intra-4f transitions of Eu3+ ions which correspond to the

    magnetic dipole transition, 5D0 →7F1 was blue shifted compared to previous result

    (Gajbhiye et al 2008 and Patra et al 1999). In europium, the (5D0 →7F1) transition

    is mainly magnetically allowed (magnetic–dipole transition) while (5D0→7F2) is a

    hypersensitive forced electric– dipole transition being allowed only at low

    symmetries with no inversion center. The electronic dipole transition, 5D0→7F2 is

    hypersensitive to Eu3+ symmetry. It could be defined by asymmetric ratio (A21) of

    the integrated intensities of 5D0 →7F2 to 5D0 →7F1 and is found to be 3.6. Lei et al

  • 79

    (2005) related the PL emissions of CdS nanostructures into band-edge and surface

    defects. Because of the quantum confinement effect, the PL peak positions of the

    band-edge emission can be related with the size of the CdS crystallites in the

    wavelength range 350–500 nm. The surface-defect emission was caused by

    surface states, such as sulfur vacancies and/or sulfur dangling bonds created by

    doping, in the wavelength range 500–700 nm. Blue emission at 430 nm was

    clearly observed which is in agreement with the earlier report (Di et al 2009) for

    CdS nanostructures. The broad emission peak between 450–480 nm, centered at 475 nm with a maximum intensity was due to the band-band and lattice defect

    emission respectively due to the CdS host and was blue shifted compared to the

    earlier reports (Savchuk et al 2006, Chahbouna et al 2007 and Cheng et al 2006).

    These observations indicate that a part of Eu3+ ions, contained in CdS nanocrystals

    and the energy can be significantly transferred from the host CdS to Eu3+ ions

    (Hayakawa et al 1999 and Okamoto et al 2002). Upon excitation, the energy from

    non-radiative recombination of electron-hole pairs can be transferred to the high

    energy levels of the Eu ions (Hayakawa et al 2000 and Reisfeld et al 2000). The

    mechanism of the intensification of (rare earth emission) REE has already been

    reported (Reisfeld et al 2000), which also supports the observed results of the

    present study. These results indicate that the adsorbed CdS particles significantly

    influence the excitation of 4f electrons in rare-earth ion. A significant change in the

    intensity of the emission bands was observed for the samples prepared using

    different mixed solvents, which are due to the variation in the structures for

    different mixed solvents. It may also be attributed to the change in the energy

    transfer rate with particle size variation and the shape of the nanocrystals.

    4.3 Ce3+

    DOPED CdS NANOPARTICLES

    4.3.1 Materials Synthesis

    The raw materials used for the synthesis of cerium doped cadmium

    sulfide [Cd1−xCexS with x = 0.01, 0.02 and 0.03] were cadmium acetate

  • 80

    20 ml solution of 0.07 M

    cadmium acetate in en+DI

    water mixed solvent

    20 ml solution of 0.07 M

    cerium nitrate in mixed

    solvent

    Mixture stirring at

    room temperature for

    30 min

    Dropwise addition of 20 ml

    solution of 0.07 M sodium

    sulphide in en+DI water.

    washed with DI water and ethanol

    Separation of nanoparticles

    by centrifugation at 4000

    rpm

    Dried in a hot air oven at 60 C

    Ultrasonication for 60

    minutes

    20 ml solution of 0.07 M

    cadmium acetate in en+DI

    water mixed solvent

    20 ml solution of 0.07 M

    cerium nitrate in mixed

    solvent

    Mixture stirring at

    room temperature for

    30 min

    Dropwise addition of 20 ml

    solution of 0.07 M sodium

    sulphide in en+DI water.

    washed with DI water and ethanol

    Separation of nanoparticles

    by centrifugation at 4000

    rpm

    Dried in a hot air oven at 60 C

    Ultrasonication for 60

    minutes

    dihydrate (99.99%), cerium nitrate (99.99%) and sodium sulfide (99.9%)

    purchased from Alfa Aesar. Mixed solvents of Ethylene diamine (EDA) and DI

    water (resistivity 10−18 mΩ) with 1:1 ratio, were used as a solvent.

    20 ml solution of cadmium acetate (99 mol. %) and 20 ml solution of

    cerium nitrate (1 mol. %) with 0.07 M were prepared separately with EDA+DI

    water as a mixed solvent. Solution of cerium nitrate was added slowly into the

    cadmium acetate solution and the mixture was subjected to stirring at the room

    temperature for 30 min. to achieve the dispersion of Cd2+ and Ce3+ into the solvent.

    With this, 20 ml solution of 0.07 M sodium sulfide was added dropwise and it was

    observed that the mixed solution turned to pale yellowish white and then to

    orange. The obtained precipitate was then ultrasonicated for 1 h to make

    nanoparticles evenly dispersed in the solution. The solution was subjected to

    centrifugation (4000 rpm) to settle down the nanoparticles and washed with

    ethanol and DI water for three times and dried in a hot air oven. Similar procedure

    was followed for other two concentrations of cerium (2 and 3 mol. %), which

    schematically shown below.

    4.3.2 Results and Discussion

    4.3.2.1 X-ray Diffraction Analysis

    The XRD patterns of pure and cerium doped CdS (Figure 4.6) show

    that the synthesized nanoparticles possess hexagonal (wurtzite) structure and the

    planes (100), (002), (101), (102), (110), (103) and (112) are clearly indexed. The

    absence of additional peaks in the XRD patterns shows the purity of the sample

  • 81

    without noticeable traces of impurity, even with increasing doping concentration.

    The prominent peaks in the diffraction pattern are (100)h, (002)h, (101)h of CdS

    (JCPDS No. 75-1545). There are no such peaks of Ce, CeS, Ce2S3 detected, thus

    confirmed the successful incorporation of Ce3+ ions into the crystal lattice of CdS

    nanoparticles. Compared to pure CdS, the diffraction peaks shift slightly toward

    smaller diffraction angle. The average crystallite size of 3.3 nm, 2.9 nm, 3.1 nm

    was calculated using Scherrer equation respectively for 1%, 2% and 3% Ce:CdS

    nanoparticles.

    Figure 4.6 XRD pattern for pure and cerium doped CdS nanoparticles

    4.3.2.2 Structural and Compositional Analysis

    TEM micrograph in Figure 4.7 (a and c) shows well-formed

    nanocrystallites of cerium doped cadmium sulfide with nearly spherical in

    shape. It clearly shows the particles within a diameter of 4 nm, which is

    consistent with the value calculated from XRD analysis. The smallest well

    dispersed particles observed from high resolution TEM image in

  • 82

    (a)

    20 nm

    (b)

    5 nm

    (c)

    50 nm

    (d)

    20 nm

    (e)

    10 nm

    (f)

    10 nm

    Figure 4.7 (b) (indicated by circles) of these ordered entities are in the

    nanoscale regime. HRTEM images of Ce doped CdS nanoparticles are also

    shown in Figure 4.7 (d-f).

    Figure 4.7 (a and c) TEM and (b & d-f) HRTEM images of cerium

    doped cadmium sulfide nanoparticles

  • 83

    EDX spectrum of Ce:CdS nanoparticles in Figure 4.8 shows the

    presence of major chemical elements namely cadmium, sulfur and cerium.

    The elemental composition calculated using ICP-OES analysis for 3% Ce

    doped CdS was 76.30, 21.15 and 2.54% of cadmium, sulfur and cerium

    respectively, which also confirms that majority of cerium ions doped with

    cadmium sulfide.

    Figure 4.8 EDX spectrum of cerium doped cadmium sulfide

    nanoparticles

    4.3.2.3 UV-Reflectance Studies

    UV-reflectance spectra of Ce doped CdS nanoparticles are shown

    in Figure 4.9. Based on these reflectance spectra the absorption edge for each

    compound was determined. Using these absorption edge values, the bandgap

    energy was estimated by extrapolating the linear region of the plot. Band-to-band

    absorption at 453 (2.73 eV), 500 (2.48 eV) and 507 nm (2.44 eV) respectively

    for 1%, 2% and 3% Ce:CdS nanoparticles shows a blue shift in comparison

    with the bulk CdS, which may be ascribed to the quantum confinement effect.

    It can also be seen that the increase in dopant percentage determines the red

  • 84

    shift in the absorption shoulder. This was clearly due to the substitution of

    Ce3+ (1.034 Å) ions into the Cd2+ (0.97 Å) which increase the size of the

    crystal lattice. This is in compliance with Brus equation (Brus 1986).

    Figure 4.9 UV-Reflectance spectra of cerium doped CdS nanoparticles

    4.3.2.4 Photoluminescence Studies

    Upon excitation at 350 nm, the samples show luminescence in the

    blue region, with an emission peak positioned around 350–450 nm as shown

    in Figure 4.10. There is no defect related emission peaks observed in the

    spectra. As shown in the inset of Figure 4.10 the peak position of NBE

    emission slightly shifts toward longer wavelength region with higher

    intensity, from 363 nm to 375 nm when the concentration of dopant increases

    to 3 mol. %. Thus it confirms that luminescence property of CdS nanoparticles

    enhanced when Ce3+ was introduced into the CdS. Maleki et al (2007),

  • 85

    attributed the peak at 363 nm, to higher level excitonic transition and its

    energy was calculated to be 3.38 eV. Similar explanation was given by

    Devi et al (2008) for the peak at 376 nm in nanocrystalline CdS. Usually the

    fluorescence emission of doping ions has higher photostability than the defect

    related luminescence of semiconductive nanomaterials, because the defects

    are greatly affected by synthesis conditions and environments.

    Figure 4.10 Photoluminescence spectra of Ce:CdS nanoparticles excited

    at 350 nm

    4.4 Sm3+

    DOPED CdS NANOPARTICLES

    4.4.1 Materials Synthesis

    In the present work, CdS nanocrystals was synthesised by chemical

    co-precipitation method using cadmium acetate dihydrate (Cd(CH3COO)2).

    2H2O, samarium nitrate hexahydrate (Sm(NO3)3. 6H2O) and sodium sulfide

    (Na2S) as Cd, Sm and S sources respectively. Samples were prepared with

    2 and 5 mol. % of samarium.

  • 86

    For 2 mol. % doping, aqueous solution of 1.305 g of 0.1 M

    Cd(CH3COO)2. 2H2O and 0.044 g of 0.1 M of Sm(NO3)3. 6H2O was

    prepared. Aqueous 0.1 M Na2S solution was also prepared separately. Na2S

    solution was then added slowly into the mixture of samarium nitrate and

    cadmium acetate solution with constant stirring to attain the orange yellow

    solution. pH of the solution was maintained at ~11 by adding aqueous

    ammonia. The solution was then refluxed with constant stirring at ~70 C for 60 min. to get a saturated solution containing Sm3+:CdS nanocrystals (NCs).

    The precipitate was ultrasonicated for 60 min. to avoid the aggregation of

    nanoparticles. It was further washed with distilled water and ethanol for

    several times to remove the organic residues present in the NCs and the

    sample was collected by centrifuging with 4000 rpm and then dried in a hot

    air oven. For 5 mol. % of doping, 0.111 g of Sm(NO3)3. 6H2O and 1.265 g of

    Cd(CH3COO)2. 2H2O were used with the same procedure discussed above.

    4.4.2 Results and Discussion

    4.4.2.1 X-ray Diffraction Analysis

    The representative XRD spectra of the Sm3+ doped CdS

    nanoparticles samples with two different concentrations are shown in

    Figure 4.11. The diffraction profile reveals that the synthesised nanocrystals

    show cubic zinc blende (-CdS) structure of CdS (JCPDS No. 10-0454). No diffraction peaks of any other minerals were detected. The crystallite size was

    calculated from a single diffraction peak using the Scherrer’s equation applied

    to the (111) reflection of cubic CdS. The corresponding crystallite size was

    estimated to be 2.62 and 2.64 nm for 2% and 5% of Sm doped CdS

    respectively.

    It was observed that the (111) peak shifted towards the higher angle

    from 26.55 to 26.85 for 2% and 26.90 for 5% Sm doped CdS nanoparticles

  • 87

    compared to bulk CdS. It reveals that the inclusion of dopant (Sm3+) ions

    substitute the interstitial sites of CdS nanoparticles to alter the crystal

    structure and lattice contraction takes place. The decrease in the lattice

    constant is due to the substitution smaller Sm3+ (0.964 Å) ions in Cd2+

    (0.97 Å) site. In addition, the broadness of the peak implies the synthesised

    Sm doped CdS particles have smaller diameter in the nanoscale regime.

    Figure 4.11 X-ray diffraction of samarium doped cadmium sulfide

    nanocrystals

    4.4.2.2 Structural, Morphological and Composition Analysis

    FESEM images of the synthesized Sm3+ doped CdS nanocrystals

    with two different magnifications are shown in Figure 4.12 (a, b). The

    particles possess almost a spherical shape with little agglomeration and the

    particles are distributed homogeneously.

    For compositional analysis, the synthesised samples were studied

    by EDX. For samarium doped CdS nanocrystals (Figure 4.13), the spectrum

  • 88

    (a) (b)

    indicates the presence of Sm, Cd and S elements and was confirmed by the

    elemental mapping shown in Figure 4.14. These results clearly reveal that the

    material synthesised are exactly Sm:CdS, having compositions consistent

    with the stoichiometric composition. The elemental composition of the

    synthesised products was further confirmed by ICP-OES analysis (Table. 4.2).

    From the results it was evident that the RE ions are incorporated into the CdS

    crystal lattice.

    Figure 4.12 FESEM images of Sm doped cadmium sulfide nanocrystals

    Figure 4.13 EDX spectrum of Sm doped cadmium sulfide nanocrystals

  • 89

    Figure 4.14 Elemental mapping of Sm doped cadmium sulfide

    nanocrystals

    Table 4.2 Bandgap energy, Specific surface area and calculated mole

    percentage by ICP-OES of Sm doped Cadmium Sulfide

    Nanocrystals

    Sample

    Mole percentage calculated

    by ICP-OES

    Bandgap

    energy

    (eV)

    Specific

    surface area

    (m2/g)Sm Cd S

    2% Sm:CdS 0.38 76.62 22.98 2.63 140.7

    5% Sm:CdS 0.77 76.11 23.10 2.58 132.9

    Cd S

    Sm

  • 90

    4.4.2.3 XPS Analysis

    X-ray photoelectron spectra (XPS) analysis of samarium doped

    CdS nanocrystals is shown in Figure 4.15 (a-d) for 5 mol. % Sm

    concentration. The survey spectrum (Figure 4.15 a) reveals the presence of

    characteristic peaks with binding energies as follows: C 1s (285.0 eV), O 1s

    (531.8 eV), Cd 3d5/2 (405.38 eV), Cd 3d3/2 (412.13 eV) and S 2p (major peak

    at 161.84 eV and a low intensity peak at 168.97 eV). In this case, the Sm 2p3/2

    binding energy is observed at 1084.37 eV. The binding energy of S 2p at

    161.84 eV also agrees well with binding energy with blue shift compared to

    the observed S 2p in pure CdS nanoparticles (Winkler et al 1999). The

    appearance of low intensity peak at 168.97 eV is due to oxidation of sulphur

    in nanoparticles (Wagner et al 1978).

  • 91

    Figure 4.15 XPS spectra for (a) wide energy (b) Cd 3d5/2 and Cd 3d3/2 in

    the 5 mol. % Sm doped CdS nanocrystals

  • 92

    Figure 4.15 XPS spectra for (c) S 2p3/2 and (d) Sm 3d5/2 in the 5 mol. %

    Sm doped CdS nanocrystals

  • 93

    20 nm

    (b)

    50 nm

    (a)

    4.4.2.4 HRTEM Analysis

    Figure 4.16 (a) and (b) shows the TEM images of synthesised Sm

    doped CdS nanoparticles for different magnifications. It can be seen from the

    figure that the nanoparticles are nearly spherical in shape with little

    agglomeration. Inset in Figure 4.16 (b) shows the SAED pattern of an area

    containing some nanoparticles. The SAED pattern shows a set of rings of the

    nanocrystals correspond to the (111), (220), and (311) planes of the cubic CdS

    phase, respectively and also demonstrate the crystallinity of the synthesised

    nanoparticles. This result is consistent with that of the XRD which composed

    of pure cubic phase CdS.

    Figure 4.16 (a, b) TEM images of samarium doped CdS nanocrystals.

    Inset in (b) shows SAED pattern

    The specific surface area (SBET) was calculated from the linear part

    of BET (Brunauer–Emmet–Teller) plots from the N2-adsorption-desorption

    isotherm shown in Figure 4.17 (a, b). The BET results (Table 4.2) showed

    that the high specific surface area for both the Sm doped CdS nanocrystals.

  • 94

    Figure 4.17 N2-adsorption-desorption isotherm of samarium doped

    cadmium sulfide nanocrystals

  • 95

    4.4.2.5 UV-reflectance studies

    To investigate the optical absorption properties, the diffuse

    reflectance UV spectra (DRS-UV) of Sm3+ -doped CdS in the range of 300–

    800 nm was examined and the results are shown in Figure 4.18. From these

    curves it is noticed that absorption is dominant mainly in blue region. The

    observed shift in the spectra is due to the size quantization effect according to

    which the bandgap value increases with the size reduction of crystallites. The

    band gap energy of this RE doped CdS nanocrystals was estimated and

    tabulated in Table 4.2. From the reflectance spectra, the absorption band edge

    was found to be 471 nm and 480 nm for 2% and 5% of Sm doped CdS

    nanocrystals respectively. While increasing the concentration, we can clearly

    observe the obvious red shift in the spectra for 5% Sm:CdS, with decreasing

    bandgap energy. Increase in the dopant concentration from 2% to 5%, shifts

    the absorption shoulder towards higher wavelength. This shift is due to the

    variation in the ionic radii of Sm3+ (0.964 Å) compared to that of Cd (0.97 Å)

    where it is substituted. This is in compliance with Brus equation (Brus 1986)

    which states that bandgap decreases with increasing size or that optical

    absorption wavelength increases with increasing ionic size.

    Figure 4.18 Reflectance UV spectra of Sm doped CdS nanocrystals

  • 96

    4.4.2.6 Photoluminescence Studies

    Figure 4.19 (a, b) shows the room temperature photoluminescence

    (PL) spectra of 2% and 5% Sm doped CdS nanocrystals respectively, obtained

    with an excitation wavelength of 350 nm. It is observed that the emission

    spectra for both the samples illustrate a band-edge (NBE) emission at ~394

    nm. It was reported that this kind of band-edge luminescence can be caused

    by the recombination of excitons and/or shallowly trapped electron–hole pairs

    (Wang et al 2001) and it was markedly blue shifted relative to that of the pure

    CdS as shown in Figure 4.19 (a). Bulk CdS reported to have a maximum

    broad emission in the 500–700 nm region of the luminescence spectra

    (Sreekumari et al 2001). For 2% of Sm doped CdS NCs the emission spectra

    in Figure 4.19 (a) includes two bands, a narrow band peaked at ~394 nm and a

    low intense peak around ~420 nm. The 394 nm emission band is attributed to

    the intrinsic emission of CdS NCs (or the band-to band transition in CdS

    nanocrystals), while the other band centered at 420 nm is attributed to the

    emission of surface states due to the small size of nanocrystals.

    Figure 4.19 (a) Photoluminescence spectra for 2% of Sm doped CdS

    nanocrystals. Inset shows a PL spectrum of pure CdS

  • 97

    Figure 4.19 (b) Photoluminescence spectra for 5% of Sm doped CdS

    nanocrystals

    It is also observed that there is a slight shift in the band edge

    emission towards higher wavelength (~398 nm) for 5% Sm doped CdS NCs.

    With increasing concentration of Sm3+ ions (5%), the peak at ~ 420 nm gets

    shifted towards longer wavelength of 524 nm. The origin of this emission

    peak, in the presence of RE ions is related to the transitions between the

    excitonic levels of CdS and the energy levels due to Sm3+, which cause shift

    in the emission peaks towards longer wavelength side. Increased in the

    concentration of RE ions improve the peak intensities, which is due to the

    energy transfer from energy levels of RE ion. Mathew et al (2008) described

    the enhancement in emission spectra is attributed to the energy transfer from

    the modified CdTe phase to the rare earth ion. Similarly Agrawal et al (2011)

    reported that, increasing La concentrations on CdS–Se films shifts the

    emission spectra towards longer wavelength, which is caused due to the

    transfer of energy from energy levels of La ion.

  • 98

    The obtained result is consistent with variations observed in the

    UV-reflectance spectra and specific surface area in the increasing RE dopant

    ions in CdS nanocrystals. When a higher percentage of dopant is

    introduced into the CdS crystal, changes in the band gap and lattice

    distortions are expected to enhance the red shift in the emission spectra

    (Shafiq and Sharif 2009).

    4.4.2.7 Raman Studies

    Micro Raman spectra for 2 and 5% samarium doped CdS

    nanocrystals are shown in Figure 4.20. The 1LO and 2LO phonon modes of

    CdS, located around 294 and 595 cm−1 respectively, are clearly identified in

    the Raman spectra of both the samples, illustrating characteristic Raman shifts

    analogous to those of pure crystalline CdS (Suh and Lee 1997). The observed

    phonon peaks are shifted toward lower frequency as expected for bulk, likely

    due to effect of small size and high surface area.

    Figure 4.20 Raman spectra of the samarium doped cadmium sulfide

    nanocrystals

  • 99

    4.5 CONCLUSION

    Eu-doped CdS nanostructures were synthesised by co-precipitation

    method using mixed solvent system. The structural and morphology changes

    in the samples prepared using different mixed solvents were studied by

    powder XRD and FESEM analyses. Samples prepared with EG+DI water

    exhibits hexagonal wurtzite phase, whereas with IPA+DI water shows cubic

    zinc-blende phase of CdS. Flower like morphology was observed for

    Eu:CdS-1 (EG+DI water) whereas clusters of rice like morphology was

    obtained for Eu:CdS-2. It was observed that the structure, crystallite size and

    the band gap of the europium doped CdS nanocrystals can be finely controlled

    by simply varying the viscosity and density of the mixed solvents. The

    incorporation of europium in the CdS lattice was confirmed by ICP analysis.

    UV-reflectance spectra reveal the blue-shift, confirming the size quantisation

    effect in the synthesized samples. In the characteristic room temperature PL

    emission spectra of the Eu3+-doped CdS, besides the peaks due to intra-4f

    transitions of Eu3+, emission peak due to CdS nanocrystals was also observed.

    The results confirm that at least a part of the Eu3+ ions is effectively doped

    into CdS nanocrystals and the energy transfer occurs from CdS nanocrystals

    to Eu3+ ions.

    Ce:CdS nanoparticles with different cerium concentrations (1, 2

    and 3 mol. %) have been synthesized by chemical co-precipitation method.

    XRD pattern reveals the formation of hexagonal structure of Ce:CdS with

    shift toward lower angle which confirms the incorporation of cerium.

    HRTEM images and SAED pattern adequately demonstrated the well

    dispersed and crystalline nature of the synthesized nanoparticles. Compared to

    pure CdS, the calculated bandgap energy value of Ce doped CdS

    nanoparticles illustrates quantum confinement effect. For 3 mol. % of cerium

    concentration, the emission intensity was found to be maximum with a slight

  • 100

    shift towards higher wavelength compared to 1 and 2 mol. %. Intense

    emission was observed due the incorporation of cerium ions into CdS NPs,

    which will be useful for high efficient EL devices and also possible for

    making full color device applications.

    Sm3+ doped CdS nanocrystals have been prepared by chemical

    co-precipitation method. The synthesised materials exhibit cubic zinc blende

    phase with an average crystallite size of ~3 nm. FESEM images show the

    aggregated morphology of the Sm doped CdS nanocrystals with high surface

    area of ~ 140 m2/g. Variation in the bandgap with doping concentrations of

    Sm3+ on CdS nanocrystals was observed using reflectance spectra. It was also

    revealed that increasing dopant concentration decreases the bandgap value of

    synthesised nanocrystals. Enhanced emission can be attributed to the energy

    transfer from CdS crystallites to RE3+ ions. Peak shift in the PL spectra are

    consistent with the successful incorporation of Sm3+ in the CdS nanocrystals.

    Front pages.docAll chapters(1.3).docCHAPTER 4-VITAE.doc