Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Nanocrystalline Titanium
Nanocrystalline Titanium
Nanocrystalline Titanium
Ebook500 pages4 hours

Nanocrystalline Titanium

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Nanocrystalline Titanium discusses the features of nanocrystalline titanium production by various SPD methods, also comparing their microstructure and properties. The authors characterize the physical, chemical and mechanical properties of ultrafine grained titanium, indicating which are crucial for their application. Titanium alloys are characterized by high specific strength combined with excellent corrosion resistance, whereas the mechanical properties of pure (or commercial purity - CP) titanium are much lower. SPD methods are proving to be an effective way to increase strength, even to a level typical for structural titanium alloys. This book is useful for academics and professionals studying the behavior of metallic materials.

  • Discusses various SPD techniques and their applications for titanium
  • Previews the limitations of SPD methods for titanium, along with the problems that can be encountered during production
  • Characterizes the physical, chemical and mechanical properties of ultrafine grained titanium and indicates which are crucial for its production applications
LanguageEnglish
Release dateSep 17, 2018
ISBN9780128146002
Nanocrystalline Titanium

Related to Nanocrystalline Titanium

Titles in the series (97)

View More

Related ebooks

Technology & Engineering For You

View More

Related articles

Reviews for Nanocrystalline Titanium

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Nanocrystalline Titanium - Halina Garbacz

    stress)

    Section 1

    Production of nanocrystalline titanium by large or severe plastic deformation

    Chapter 1

    High-pressure torsion and equal-channel angular pressing

    Irina P. Semenova⁎; Ruslan Z. Valiev⁎,†; Terence G. Langdon‡    ⁎ Institute of Physics of Advanced Materials, Ufa State Aviation Technical University, Ufa, Russia

    † Laboratory for Mechanics of Bulk Nanostructured Materials, Saint Petersburg State University, Saint Petersburg, Russia

    ‡ Materials Research Group, Faculty of Engineering and the Environment, University of Southampton, Southampton, United Kingdom

    Abstract

    In this chapter, the main peculiarities of ultrafine-grained (UFG) structure formation in commercially pure Ti is examined when processing by high-pressure torsion (HPT) or equal-channel angular pressing (ECAP). High-pressure torsion is used to produce Ti with extremely small grain sizes (generally < 100 nm). The ECAP technique was developed in the 1990s by R.Z. Valiev and coauthors and successfully used to produce UFG structures in bulk billets of pure Ti. Recently, this technique was further developed and a number of modifications of the technique have been proposed, including continuous ECAP or ECAP-Conform which enables the production of long-length rods (up to 1 m).

    Keywords

    Titanium; Severe plastic deformation; Nanostructuring; Ultrafine-grained structure

    Acknowledgments

    The work of R.Z.V. was supported in part by Saint Petersburg State University in the framework of Call 3 project (id 26130576) and in part by the RF Ministry for Education and Science under Grant agreement No. 14.586.21.0061 (unique project identifier RFMEFI58618X0061). I.P.S. thanks the Ministry of Education and Science of the Russian Federation within the scopes of the basic part of the state assignment for educational institutions (Grant No. 16.5979.2017).

    1.1 Nanostructuring of titanium by high-pressure torsion

    Studies performed over the last decades have demonstrated that the formation of the nanostructured state in a material is an efficient way to improve its physical and mechanical properties [1–6]. One of main approaches to producing nanostructure materials is based on refining the microstructure until nanograins in bulk samples using severe plastic deformation (SPD) methods [3–6]. Its principle is that during SPD materials are subjected to large deformations under conditions of high applied pressures [4]. As ultrafine-grained (UFG) metals and alloys fabricated using SPD methods have, as a rule, grain sizes in the range of 100–500 nm and usually contain different nanostructural elements inside the grains (nanotwins, nanoparticles, segregations, etc.), these elements significantly affect the material properties. As a result, the materials are assigned to the class of nanostructured bulk materials [6]. High-pressure torsion (HPT) and equal-channel angular pressing (ECAP) techniques were successfully applied to produce UFG structures in pure Ti. However, the structural features of UFG titanium strongly depend on the method and the treatment regime. The features of the formation of a nanostructured state in commercially pure titanium by the HPT method are discussed in this chapter.

    Specimens from commercially pure (CP) titanium VT1-0 (Russia) were initially processed by HPT under an applied pressure of 5 GPa to a true logarithmic strain of about 7 at room temperature [7]. They had the shape of thin discs with a thickness of 0.3 mm and diameter of 10 mm. Fig. 1.1 shows typical transmission electron microscopy (TEM) images of the HPT-processed structure in CP Ti at different magnifications. It is seen that a homogeneous UFG structure with a mean grain size of about 100 nm is formed (Fig. 1.1A and B). This structure is characterized by the following features: first, by extinction contours which are clearly seen in the grain interiors indicative of large internal stresses; second, grain boundaries appear to be of mostly high angle misorientation but they are not well-defined and there is no banded contrast at the boundaries that is characteristic of well-formed grain boundaries as usually observed by TEM [7,8]. The HRTEM images show that the grain boundaries in the processed material are typically wavy, curved, or corrugated (Fig. 1.1C) indicating their highly nonequilibrium character [4,9]. Moreover, a variation in the misorientation angle is observed along the same boundary by about 5 degrees, probably due to disclinations. A high level of defects is also observed in the grain interiors. In many zones of the grain interior the lattice fringes are poorly defined (area marked by the square in Fig. 1.1C) suggesting local distortions of the crystalline lattice.

    Fig. 1.1 Typical TEM bright field (with SAED as an inset ) (A) and dark-field (B) micrographs and HRTEM image (C) of as-processed CP Ti [9].

    In Ref. [9] an unusual aspect of low-temperature annealing on the boundary structure was observed. It was shown that during annealing, grain growth in the HPT-processed Ti starts at a temperature of 350°C and higher. An analysis of short-term annealing for 10 min at temperatures of 250°C and 300°C impacted the microstructure of the HPT-processed Ti and revealed important changes as compared to the as-produced state [9]. First, there is a significant decrease in the lattice distortions without visible grain growth during annealing. At a higher resolution it was shown that annealing at 250°C revealed a rearrangement of defects resulting in their movement from the grain interiors to near the grain boundary areas. An atomic resolution image of such a typical highly distorted grain boundary area is shown in Fig. 1.2A. An ordered array of dislocations [10] with periodicity of λ = 0.7 nm can be observed in some areas of the high angle grain boundary. The schematic illustration of these defect regions is presented in Fig. 1.2B. The estimated dislocation density is very high and is equal to about 10¹⁴ m− 2, which is indicative of the nonequilibrium character of the grain boundaries after this annealing. However, as with any ordering phenomenon, this process lowers the internal stresses (and, therefore, the free energy) in the structure as a system that was confirmed by an analysis of XRD patterns which exhibited peak broadening after this heat treatment [9].

    Fig. 1.2 (A) HRTEM of GB region in CP Ti after HPT and annealing at 250°C for 10 min; (B) scheme of the ordered defect structure in selected area [9].

    In the late 1990s, researchers showed that the consolidation of metal powders via SPD techniques could be implemented to prepare high density samples with grain sizes smaller than 50 nm [11,12]. In a recent study [13] there was a report of a novel processing approach for producing nanocrystalline CP Ti with grain sizes smaller than 40 nm. The results support the hypothesis that it is possible to synthesize nanostructured CP Ti of high density by using a new approach involving HPT processing of powder after cryomilling.

    Fig. 1.3 shows bright-field and dark-field TEM images of the microstructure at the peripheral zone (approximately 3 mm from the center) of the sample after HPT. This region does not show any significant porosity and the microstructure of Ti consists of ultrafine grains with an average size ranging from ~ 25 to ~ 40 nm (see Fig. 1.3A). However, some coarser grains with an average size of ~ 120 ± 50 nm are also observed in the microstructure (see Fig. 1.3B). The fraction of the large grains in the microstructure does not exceed 7% and it is likely that the large grains in the Ti sample are formed during HPT through coalescence of the initial small grains in the cryomilled Ti powders. A similar phenomenon was observed in Cu powders subjected to HPT [14]. In this study a possible mechanism of nonuniform grain growth can be also associated with dynamic recrystallization, because SPD processing is performed at elevated deformation temperatures (T ≈ 300°C) when grain boundary diffusional processes are activated. The appearance of a nonuniform contrast within the grains in the dark-field images indicates a high lattice distortion (see Fig. 1.3C). The total dislocation density in the grain interiors is estimated to be approximately 2 × 10¹⁴ m− 2. The spots on the selected area electron diffraction (SAED) patterns form concentric circles (see Fig. 1.3A) and this feature is consistent with the formation of high angle boundaries which are typical of nanostructured materials produced by HPT under the conditions that involve a high number of rotations [4]. The microstructure in the central region of the sample revealed a nonuniform distribution of grain sizes ranging from ~ 25 to ~ 240 nm (see Fig. 1.4). Such nonuniformity of the structure formed in the center of HPT samples can be attributed to the competing processes of deformation, dynamic recrystallization, and grain growth, which are likely to be activated at elevated temperatures (300°C).

    Fig. 1.3 (A, B) Bright-field and dark-field images of ultrafine grains formed in peripheral part of the sample; (C) ultrafine grains [13].

    Fig. 1.4 (A) TEM micrographs coarse grains; (B) ultrafine grains in the center of the sample [13].

    Thus, HPT of CP Ti at room temperature may result in a nanostructured state formation characterized by a very small grain size from ~ 120 to ~ 40 nm, a high dislocation density, enhanced internal stresses, and high angle misorientations of the grain boundaries.

    1.2 Grain refinement by ECAP and its modification

    The ECAP technique that involves deformation of massive samples by simple shear was developed by V.M. Segal with his colleagues in the 1970s. It subjected materials to plastic strains without changing their cross-section and therefore permitting repeated deformation [15]. The required condition for UFG structure formation in the billet is shear at the point of channel intersection at lowered temperatures (lower than the recrystallization temperature) ensuring an achievement of the accumulated strain degree of ε ~ 4–6. In the early 1990s, Valiev and coauthors developed this technique and applied it as an SPD technique to form structures with submicrocrystalline and nanometer grain sizes [4,16].

    Usually at frequently used angles of straining of 90 degrees, each pass corresponds to the additional strain degree of about 1 [17,18]. Furthermore, the billet orientation when passing through the channels or the route is an important parameter of ECAP. In Refs. [19–21] different ECAP techniques were considered: the billet orientation remained unchanged at each pass (route A); after each pass the billet was rotated about its axis by 90 degrees (route B); after each pass the billet was rotated about its longitudinal axis by 180 degrees (route С). Later, the subtypes of the route В, called ВА and ВС, were introduced for consideration. These two subtypes differ in the rotation direction between the subsequent passes. Thus, route ВА consists of alternative rotation clockwise and counterclockwise by 90 degrees and in route ВС the rotation is performed only clockwise. These routes differ in the shear direction at multiple passes of the billet through the intersecting channels.

    Thus, processing by ECAP is multifactorial. Its successful implementation depends on the die-set geometry (angle of channels intersection, their shape and sizes), the parameters of ECA pressing (velocity, number of passes, route, temperature, lubricant), and the nature of the deformed material (initial structure, ductility, strength). One should also take into account possible heating of samples during ECAP [22–24], which also affects the microstructure formation. As is known, depending on the oxygen content and other impurities, commercially pure Ti can be of Grade 2, Grade 3, or Grade 4, which can significantly impact the mechanical properties of Ti in coarse-grained and UFG states

    Enjoying the preview?
    Page 1 of 1