Diamond Growth and Morphology under the Influence of Impurity

Oct 18, 2013 - ABSTRACT: In this paper, we report on the diamond growth and morphology evolution of diamond crystals of cube-octahedron series under t...
3 downloads 0 Views 4MB Size
Article pubs.acs.org/crystal

Diamond Growth and Morphology under the Influence of Impurity Adsorption Yuri N. Palyanov,*,†,‡ Alexander F. Khokhryakov,†,‡ Yuri M. Borzdov,† and Igor N. Kupriyanov† †

Sobolev Institute of Geology and Mineralogy, Siberian Branch of Russian Academy of Sciences, Koptyug ave 3, Novosibirsk, 630090, Russia ‡ Novosibirsk State University, Novosibirsk, 630090, Russia ABSTRACT: In this paper, we report on the diamond growth and morphology evolution of diamond crystals of cube-octahedron series under the effect of impurity adsorption. In the first series of experiments it is established that in the Ni0.7Fe0.3−C system without impurity additives diamonds of predominantly cubic, cuboctahedral and octahedral habit crystallize at 1300, 1335, and 1370 °C, respectively. In the second series of experiments, the effect of H2O additive, in amount 0.3 wt %, on diamond growth and morphology was studied at different temperatures. It is found that the impurity adsorption effect leads to growth inhibition followed by complete blocking of growth of the {100} faces. At these conditions, the {111} faces continue to grow but reduce in area due to the inhibition of the ends of the {111} growth layers. As a result, a rich morphological diversity of antiskeletal diamond crystals was produced, whose growth and design are determined by the impurity adsorption effect and crystallization temperature. The established regularities of the crystal morphology evolution are in good agreement with the existing concepts of crystal growth under the impurity adsorption effect and structural crystal-chemical features of the diamond faces. Metal−carbon complexes containing hydrogen related groups together with hydroxyl, carbonyl and carboxyl groups are supposed to act as the adsorbing impurities.



{911} faces.22 In carbonate, carbonate-silicate and carbonatechloride media, modeling composition of inclusions in natural diamonds, it is established that diamond morphology is strongly altered by the addition of H2O and the only growth form at these conditions is octahedron.23,24 Thus, establishing the regularities of changes in morphology, real structure and properties of diamond with reference to the growth conditions is still an important task, since it is of interest both for the reconstruction of complex processes of diamond formation in nature,25,26 and for growth and design of crystals in laboratory. The significant growth sector inhomogeneity of diamond crystals and the need for production of diamond crystals with controlled properties, recently commenced active studies of the effect of impurities on diamond growth, morphology and properties using the most efficient metal− carbon systems.27−32 It has been found that the most dramatic changes in the diamond morphology occurs with the addition of H2O to the metal melt, which acts as a surface-active impurity and results in crystallization of diamond rhombic dodecahedrons and dendrites.29,33,34 To get further insights into the effect of impurity adsorption on growth and morphology of diamond we have undertaken an experimental study on diamond

INTRODUCTION Analysis of data on the diamond morphology1−6 shows that crystals obtained from various deposits are highly diverse, implying a considerable range of conditions for diamond formation in nature. Significant morphological diversity of natural diamonds is determined not only by the relative development of crystal faces of the octahedron, cube and dodecahedron, but also by the abundance of rounded surfaces,1,2,7 elements of the antiskeletal structure of faces,2,4 and the specific manifestations of “fibrous” growth.2,8−10 As is known natural diamonds usually show a very complex internal structure associated with the inhomogeneous distribution of defects and impurities.11−16 The morphology of the majority of synthetic diamonds is determined by the relative development of the {100} and {111} faces, which are either flat or have a skeletal structure. It is well-known that the morphology of synthetic diamond crystals synthesized in metal− carbon systems changes from cube to octahedron with increasing temperature.17,18 It is experimentally shown that the composition of the crystallization media is another important factor, controlling diamond growth and morphology. In the CaCO3−C system, diamond morphology is characterized by tetragon− trioctahedron {533}, {955}, {755}, {211}, and {322} faces.19 The only growth form of diamond in sulfide melts is octahedron.20 In the sulfur−carbon system diamond morphology is determined by the {100} and {111} faces with minor development of {411} and {944} ones.21 Diamonds crystallized in the phosphorus−carbon systems exhibit the {111}, {310}, and © 2013 American Chemical Society

Received: September 9, 2013 Revised: October 17, 2013 Published: October 18, 2013 5411

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Figure 1. SEM images of a cubic diamond crystallized under the impurity adsorption effect. (a) Overall view and (b, c, d) growth relief on the {100} faces.

diamond was completely blocked at CH2O ≥ 0.43 wt % and metastable graphite instead of diamond crystallized. In the present work to study the impurity adsorption effects and associated changes in crystal growth and morphology, the concentration of H2O additive was set 0.3 wt %. Several additional experiments were performed to estimate the rate of H2O intake into the crystallization medium and to define optimum run duration. At 1335 °C plane-faced cuboctahedral diamond crystals were produced in a 30-min run, and crystals exhibiting slight signs of growth inhibition were found in a 1-h run. Assuming that with further accumulation of H2O in the crystallization medium the growth rate would significantly decrease, run duration in the main series of experiments was set 15 h. Summarizing, experiments on diamond crystallization under the impurity adsorption influence were performed at 6 GPa and temperatures 1300, 1335, and 1370 °C with CH2O = 0.3 wt % and run duration of 15 h. For this P−T range, the accuracy of temperature and pressure measurement is ±0.1 GPa and ±20 °C. At each temperature four experiments were carried out. The diamond crystals produced were studied using optical and scanning electron microscopy. A Tescan MIRA3 LMU SEM was used to study the morphology of the crystals. A 50-Å thick Cr film was deposited on the samples using Ar-ion sputtering prior to SEM examination. Crystallographic indices of the faces were determined by a GD-1 two-circle goniometer with the uncertainty of 3−5′. Diamond crystals with complicated surfaces were studied using a photogoniometer with a cylindrical

crystallization in the Ni−Fe−C system with H2O additive at different temperatures.



EXPERIMENTAL SECTION The experiments were performed at a pressure of 6 GPa using a multianvil split-sphere apparatus. The design of the apparatus and scheme of high-pressure cell for diamond crystallization has been presented elsewhere.28 A temperature gradient growth (TGG) method was applied for diamond crystallization. No diamond seed crystals were employed to ensure spontaneous crystallization of diamond producing a large number of crystals needed for investigations. A Ni0.7Fe0.3 alloy, produced from a mixture of nickel (1900 mg) and iron (814 mg) powders, both 99.96% pure, was used as the solvent−catalyst. Graphite with purity 99.99% was used as a source of carbon. A mixture of Mg(OH)2 and SiO2, which reacted under the experimental conditions to produce Mg2SiO4 and H2O, was used as a source of water. The sample assembly was the same as in our previous study,29 except for the seed crystals, which were not added. A MgO sleeve was used to isolate the sample assembly from the heater. In a series of preliminary experiments, performed without H2O additives, the temperatures of crystallization of diamond crystals with predominantly cubic, cuboctahedral, and octahedral habit were determined to be 1300, 1335, and 1370 °C, respectively. Choosing the H2O additive concentration, we took into account our previous results,29 which showed that with an increase in H2O additive concentration (CH2O) in the Ni0.7Fe0.3−C system in the range from 0 to 0.5 wt %, the crystallization of 5412

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Figure 2. SEM images of antiskeletal diamond crystals formed from initial cube-octahedrons during impurity inhibited growth of the {100} faces. (a−d) Overall view and (e, f) growth layers on the {100} faces.

that of the {111} faces. Therefore the {100} face can be regarded as a face of inhibited growth, i.e. a face whose growth rate is almost zero at nonzero supersaturation. As a result of impurity poisoning various specific growth patterns appear on the inhibited {100} faces. As shown in Figure 1 growth layers are characterized by a pronounced serrated structure. Growth microlayers are frequently overgrown with thick macrolayers (Figure 1b). At some sites the effect of growth inhibition is manifested (Figure 1c). It is noted that macrolayers significantly increase in thickness, elements of many-headed growth appear, and specific outgrowths or protuberances with the {111} micro faces are formed (Figures 1d and 2f). The specific features of the relief of the {100} faces indicate that the sources of growth layers were active, but the propagation of the layers over the face was strongly inhibited or blocked, that is, from a certain point {100} faces were completely poisoned with impurities. At the conditions when growth over {100} faces was stopped, the {111} faces continued to grow but decreased in area. Consequently, the impurity adsorption poisons not only {100} faces, but also the ends of the growth layers of relatively fast-growing (noninhibited) {111} faces. As a result, the {111} faces formed pyramids with side surfaces having a stepped structure (Figures 1a, 2a−d) or rough relief (Figure 2e). Depending on the number of the active growth centers and the extent of growth blocking, either one growing-out pyramid, occupying the entire face (Figure 2a, d), or a series of separate pyramids (Figure 2b, c) are formed. The crystallographic orientation of the side surfaces adjacent to the {100} faces may coincide with the {hkk} faces (tetragon-trioctahedron series), {111} faces, and even with the {100} faces. According to goniometry, the tetragon-trioctahedron

photographic chamber. In this case, face coordinates were determined with the uncertainty of 30′. Infrared absorption spectra were recorded using a Bruker Vertex 70 Fourier transform infrared (FTIR) spectrometer fitted with a Hyperion 2000 microscope. Concentration of nitrogen impurity in diamond crystals was derived from the infrared spectra using standard procedures.35 Raman and photoluminescence (PL) spectra were measured using a Horiba J.Y. LabRAM HR800 microspectrometer with an Ar-ion laser (514 nm).



RESULTS Control experiments in the Ni0.7Fe0.3−C system without H2O addition were performed at 1300, 1335, and 1370 °C. At 1300 °C plane-faced crystals of cubic habit with minor or nonexistent {111} faces were produced. At 1335 °C, diamond crystallized in the form of cuboctahedron with nearly equal development of the {100} and {111} faces. Diamond crystals of octahedral habit with minor {100}, {110}, and {311} faces were produced at 1370 °C. The effects of impurity adsorption on diamond growth and morphology were found to be most prominent at 1300 and 1335 °C, corresponding to the crystallization of cubic (Figure 1) and cuboctahedral diamonds (Figure 2). Since the intake of H2O into the crystallization medium was gradual, because of both the kinetics of the H2O generating reaction and the diffusion of water through the carbon source, in the course of experiments, at first diamonds crystallized in the form of plane-faced crystals and then with increasing impurity concentration, impurity adsorption started to affect diamond growth. Typical crystals shown in Figures 1 and 2 clearly demonstrate that at a certain moment growth of the {100} faces either was stopped completely or the growth rate of the {100} faces became incomparably lower than 5413

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Figure 3. Scheme of transformation of a cubic diamond crystal with minor {111} faces during impurity inhibited growth. The {100} faces are shown by yellow, {111} faces are blue, and side surfaces of the {111} pyramids are red. This color coding also applies for Figures 4−6 and 9−11

Figure 4. Scheme of transformation of a cube-octahedral diamond crystal during impurity inhibited growth.

The scheme of transformation of such crystals is presented in Figure 6. At a temperature of 1335 °C, where cuboctahedral diamonds crystallize, specific crystals of another type were produced under the effect of impurity adsorption (Figure 7). In this case, negative tetragonal pyramids on the {100} faces and serrated growth macrolayers on the {111} faces are formed. Different stages of the development of such diamond crystals are demonstrated in Figure 7. The negative tetragonal pyramids, formed by the {111} faces can be individual (Figure 7a) or multiple (Figure 7b). It is possible that in these cases growth of the {100} faces was locally blocked by the impurity adsorption at some places on the face. At a temperature of 1370 °C, where octahedral diamonds crystallize, there are no crystal faces of inhibited growth, that is, the {100} faces. However, the form of the crystals changes due to the inhibition of the ends of the {111} growth layers from the side of edges and vertices of octahedron and the formation of growing out pyramids of the {111} faces. The main types of the crystals formed from the initial octahedrons under the effect of impurity adsorption are shown in Figure 8. At the initial stage of inhibition, octahedrons with re-entrant angles at octahedron edges and the tetragonal depressions at the octahedron vertices are formed (Figure 8a). A crystal shown in Figure 8b illustrates the next stage of the evolution of the octahedron under the impurity adsorption effect. When the {111} faces grow out completely, a

Figure 5. Schemes of the internal structure of antiskeletal diamond crystals formed from initial cube-octahedrons.

faces have {16.1.1.} and {18.1.1} indices. In the end, {111} faces grew out completely and the crystal growth ceased. The schemes of formation of such crystals are shown in Figures 3 and 4, and their internal structure is shown in Figure 5. When growth of the {100} faces is inhibited, the shape of the {111} growth pyramids is largely determined by the geometry of the octahedral faces of the initial crystals. If the {111} faces have a triangular form, then trigonal growing out pyramids are formed, as shown in schemes presented in Figures 3 and 4. When in the initial crystal habit the {111} faces dominated over the {100} ones, as a result of the inhibition of the {100} faces, ditrigonal pyramids are formed as shown in Figure 2d.

Figure 6. Scheme of transformation of a diamond octahedron with minor {100} faces during impurity inhibited growth. 5414

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Figure 7. SEM images of diamond crystals grown at the conditions of local impurity blocking of {100} faces.

impurity induced growth inhibition, represented by the dark heavily included rim and morphology modification, infrared measurements gave as a rule strongly distorted absorption spectra. Nevertheless, in most cases it was possible to identify absorption due to the nitrogen impurity (a band peaking at 1130 cm−1) and absorption due to presumably C−H stretching vibrations of hydrocarbons (a band at 2840 cm−1) (Figure 11b). Raman spectra recorded for these diamonds show significantly enhanced photoluminescence from the nitrogen-vacancy centers (Figure 12b). In addition a weak band at around 1550 cm−1 can be seen in the spectra. This feature can tentatively be assigned to the G band of highly disordered graphite, which could probably form upon dissociation of hydrocarbons or upon quenching from carbon dissolved in fluid inclusions.

polyhedron resembling a penetration twin of two tetrahedrons (Figure 8d) is formed. The general scheme of the transformation of the octahedron under the adsorption effect of impurities is shown in Figure 9. On the real crystals (Figure 8) side surfaces of the {111} growing out pyramids can be flat, concave or convex. With a constant slope and the absence of rough relief, these surfaces may correspond to the {111} and {110} faces or {hhl} faces of a trigontrioctahedron series, among which {551}, {881}, and {553} faces are identified by the goniometry measurements. In all cases, these faces are “false” or “passive” since they do not have corresponding growth pyramids. Schemes illustrating the internal structure of such crystals are presented in Figure10. As already noted, when the growing out pyramids of {111} faces are formed, the ends of the inhibited growth layers may have different slopes. As a result, some growth layers grow beyond the pyramid and form specific outgrowths or protuberances (Figures 2a, d and 8b, d). For a number of diamond crystals selected from the experiments, infrared absorption and Raman measurements were performed. From the infrared measurements it is found that diamond crystals produced in the control experiments and those crystallized in the main series of experiments, but at the conditions of negligible effects of impurity action, show infrared spectra characteristic of type Ib diamonds and contain single substitutional nitrogen in concentration 200−250 ppm (Figure 11a). Raman spectra of such diamonds demonstrate an intense peak at the diamond Raman frequency 1332 cm−1 and, occasionally, a relatively weak peak at about 1.945 eV (638 nm), which is due photoluminescence from negatively charged nitrogen-vacancy centers (Figure 12a). For the crystals exhibiting features of



DISCUSSION The experimental data on diamond crystallization in the metal− carbon melt in the presence of H2O additives, obtained in this and previous29,33,34 studies allow us to consider the established regularities of growth and significant changes of morphology of diamond crystals as a phenomenon caused by the adsorption effect of the impurities. According to the current concepts,36,37 impurity species adsorb on the surface of the growing crystal, and the attending processes can be quantitatively described by the mathematical apparatus of the theory of adsorption. One of the first models aimed at explaining the empirical fact of the existence of the dead zones in the presence of an impurity component, depending on the supersaturation, is the model of Cabrera and Vermilyea.38 According to this model the impurity strongly 5415

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Figure 8. SEM images of antiskeletal diamond crystals formed from initial octahedrons during impurity inhibited growth.

Figure 9. Scheme of transformation of an octahedral diamond crystal during impurity inhibited growth.

increasing impurity concentration. The micromorphology of faces under the influence of impurities changes toward coarsening of relief: the thickness and serration of growth layers increases. Deceleration of the growth rate of faces or complete blocking of face growth may lead to significant changes in the crystal morphology.41,42 However, experimental studies of such morphological transformations are still rare. It should also be noted that the impurity adsorption effect and the growth inhibition of different faces will be determined not only by the growth conditions, but also by the structural features of the faces. From the structural and crystal-chemistry features of the main faces of diamond and the periodic bond chain (PBC) analysis,3,43,44 it follows that for materials with the diamond structure, the {111} faces corresponds to F type, the {110} and trigon-trioctahedral faces to S type, and the rest, including the {100} and tetragon-trioctahedral faces to K type. In general one can expect that the impurity induced growth inhibition is less

adsorbs on the terrace and ledges between the growth layers and prevents their tangential propagation. The Chernov−Bliznakov model,36,39 considers the impurity adsorption at growth step edges and kinks, but does not assume a complete blocking of face growth over a broad range of supersaturation and concentration of impurities. The possibility of impurity adsorption on the surface with the formation of a two-dimensional adsorption layer has also been considered.37 The unequivocal selection of the model of the impurity adsorption and induced growth inhibition is complicated by the fact that in the real systems the results depend on the thermodynamic, kinetic, and many other factors,36,37,40 which are difficult to take all into account even for the crystallization of simple materials from aqueous solutions. Considering the existing models of the impurity adsorption on the growing crystal surfaces, and empirical data on the growth of crystals it can be noted that the minimum value of supersaturation sufficient for crystal growth generally increases with 5416

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

and changes in diamond properties in these growth zones. However, in the case of simple deceleration of growth rate of {100} faces and, consequently, the change of the ratio of growth rates of the {100} and {111} faces, diamond crystals with dominant {100} faces should ultimately form. The crystals with protrusive {111} faces, which were produced in our experiments, can be derived from the polyhedrons of the cube-octahedron series only in the case of a complete blocking of the {100} faces by adsorbed impurities and, accordingly, complete cessation of their growth starting from a certain moment is assumed. The general scheme of the morphological changes of diamond crystals with reference to the crystallization temperature and the impurity adsorption effect is shown in Figure 13. In fact, it represents the transformation scheme of plane-faced crystals of the cube-octahedron series into antiskeletal polyhedrons. The antiskeletal structure of {111} faces is realized upon the inhibition and subsequent blocking of growth of {100} faces, when the only growing faces are the octahedral faces. An important aspect in the development and design of these crystals is that in addition to the blocking of {100} faces, the ends of the {111} growth layers, adjacent to the {111} faces, are also affected by the impurity poisoning. In general, the morphology of antiskeletal diamond crystals is determined by the habit of initial crystals, the structure of the side surfaces of {111} growth pyramids, the degree of truncation of the pyramids and the symmetry of their development. It should be noted that the polyhedrons shown in Figure 13 are idealized and do not reflect the diversity of the observed shapes of antiskeletal diamond crystals, primarily because of different inclination angles of the side surfaces of the {111} pyramids, possible significant variations of these angles for real crystals and their symmetry. The question of the composition of the impurities which adsorb on diamond crystal surfaces and lead to the morphological changes remains unclear. It has been established experimentally that the primary cause for these phenomena is H2O impurity, resulting from the reaction between Mg(OH)2 and SiO2. Previously we showed that the f O2 value in experiments on diamond crystallization in reduced metal−carbon systems with the addition of H2O, at the first approximation corresponds to the buffer IW (iron/wustite).29 This inference is based on the fact than an increase in H2O content led to enlargement of a zone of magnesiowustite, which was formed at the contact between the Fe−Ni melt and the isolating sleeve made of MgO. Thermodynamic calculations show that at 6.0 GPa, 1400 °C, and f O2 near the IW buffer, the equilibrium C−O−H fluid consists of methane (55 mol %), water (42 mol %) and hydrogen (3 mol %)45 and may also contain other hydrocarbons, e.g. C2H6.46 Experimental data suggest that at similar P, T, and f O2 conditions methane can possibly dissociate with the formation hydrogen, carbon and higher hydrocarbons.46,47 However, under the crystallization conditions, not the fluid components but the products of their reactions, first of all with metal melt and carbon should be considered as the adsorbing impurity. Multiple inclusions detected in the outer zones of the crystals are represented by metal−carbon melt and metal-free fluids of hydrocarbon composition, such as those previously established in diamonds of the rhombic dodecahedron habit.29,48 This indicates that the crystallization medium was heterogeneous consisting of metal melt with dissolved fluid components and segregated micro bubbles of mainly hydrocarbon composition. Segregation of hydrocarbons implies limited and low solubility of

Figure 10. Schemes of the internal structure of antiskeletal diamond crystals formed from initial octahedrons. Side surfaces of the growing out pyramids of {111} faces correspond to (a) rhombic dodecahedral and (b) octahedral planes.

Figure 11. Typical infrared spectra recorded for diamond crystals produced as a result of (a) noninhibited and (b) impurity inhibited growth. The spectra are displaced vertically for clarity.

Figure 12. Typical Raman/PL spectra recorded for diamond crystals produced as a result of (a) noninhibited and (b) impurity inhibited growth. The spectra are displaced vertically for clarity.

likely for F type faces, that is, {111} faces, and most likely for K type faces, that is, {100} faces. Indeed on a number of real diamond crystals of cuboctahedral habit the inhibition of growth of the {100} faces is observed, which is manifested in the increase in growth layer thickness and serration. As a rule, the stage of growth inhibition (growth rate deceleration) is accompanied by intense entrapment of inclusions of the crystallization medium 5417

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Figure 13. General scheme of the effects of crystallization temperature and impurity adsorption on diamond crystal habit.

natural and synthetic diamond are very likely similar and related to the adsorption effect of impurities.

hydrocarbons in the metal melt and may lead to an increase in the molar ratio of H2O fluid dissolved in the melt. Discussing the question on the possible nature of the adsorbing impurities is of interest to consider the results of high resolution X-ray photoelectron spectroscopy,49,50 which show that in hydrogen atmosphere, hydrogen related structures, such as the monohydride (−CH), dihydride (−CH2), and trihydride (−CH3) are formed on the diamond surface. Upon subsequent wet chemical oxidation treatment a conversion from hydrogenated to oxygenated surface takes place and hydroxyl (C−OH), carbonyl (CO) and carboxyl (OC−OH) groups are formed on the diamond surface.50 For our discussion, two points, which follow from these studies, worth noting. First, the oxidation level on diamond (001) surface is much heavier than on (111) surface, and second, the hydroxyl (C−OH) groups are the most abundant on the (100) and (111) surfaces of chemically oxidized diamonds. Summarizing the above considerations, we suppose that the adsorbing impurities in our experiments can be metal−carbon complexes containing in its structure hydrogen related groups together with hydroxyl (C−OH), carbonyl (CO), and carboxyl (OC−OH) groups. Analysis of the existing data on the morphology of natural diamonds1−5 and comparison with the crystals produced in this study show that the antiskeletal polyhedrons of synthetic diamond grown under the impurity adsorption influence are to a great extent morphological analogues of some natural diamonds. For example, diamonds with a plate-stepped structure of the octahedral faces, forming combination surfaces are frequently found in nature. Octahedral crystals with re-entrant angles at the edges and crystals resembling penetration twins of tetrahedrons are also well-known among natural diamonds.1,2 Natural diamond crystals are very diverse and have, as a rule, very complex structure,2,4,5,9,51−55 but even in these cases, there are significant similarities with those polyhedrons obtained in this study. On the basis of the morphological similarity, we can conclude that the processes determining the morphology of antiskeletal crystals of



SUMMARY (1) In a series of experiments on diamond crystallization in the metal−carbon system with H2O additives the formation of specific antiskeletal diamond crystals is established. The growth and design of these crystals is controlled by the adsorption effect of impurities and depend on the crystallization temperature. (2) Plane-faced diamond crystal of the cuboctahedron series evolve into antiskeletal crystals as a result of blocking of the {100} faces and inhibition of the ends of the {111} growth layers induced by impurity poisoning. The impurity adsorption leads to the formation of convex polyhedrons, whose morphology is determined by the habit of the initial crystals, the structure of the side surfaces of the {111} growth pyramids, the degree of truncation of the pyramids and the symmetry of their development. (3) The established regular changes in the diamond morphology are in good agreement with the existing concepts on crystal growth under the adsorption effect of impurities and with structural and crystal-chemical features of the diamond faces (PBC analysis). Metal−carbon complexes comprising hydrogen related groups together with hydroxyl, carbonyl and carboxyl groups are supposed to act as the adsorbing impurities



AUTHOR INFORMATION

Corresponding Author

*Tel./Fax: +7-383-3307501. E-mail: [email protected]. Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding

Support from the Presidium Program of the Russian Academy of Sciences (grant no. 2.1) is gratefully acknowledged. 5418

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419

Crystal Growth & Design

Article

Notes

(35) Zaitsev, A. M. In Handbook of Industrial Diamonds and Diamond Films; Prelas, M., Popovici, G., Bigelow, L., Eds.; Marcel Dekker Inc: New York, 1997, pp 227−376. (36) Chernov, A. A. Modern Crystallography III. Crystal Growth; Springer-Verlag: Berlin, 1984; pp 104−204. (37) Sangwall, K. Etching of Crystals: Theory, Experiment and Application; Mir: Moscow, 1990; p 492 (in Russian). (38) Cabrera, N.; Vermilyea, D. A. In Growth and Perfection of Crystals; Doremus, R. H., Roberts, B. W., Turnbull, D., Eds; John Wiley: New York, 1958; pp 393−408. (39) Bliznakov, G. M. Sov. Phys. Crystallogr. 1959, 4 (3), 134−140. (40) Land, T. A.; Martin, T. L.; Potapenko, S.; Palmore, G. T.; De Yoreo, J. J. Nature 1999, 399, 442−445. (41) Punin, Yu. O.; Petrov, T. G.; Treyvus, E. B. Zap. Ross. Mineral. Ova. 1980, 109, 517−529 (in Russian). (42) Treivus, E. B. The Kinetics of Crystal Growth and Dissolution; Leningrad University: Leningrad, 1979; p 248 (in Russian). (43) Hartman, P. Z. Kristallogr. 1963, 119, 65−78. (44) Sunagawa, I. In Handbook of Crystal Growth, Vol. 2; Hurl, D. T. J., Ed.; Elsevier Science: North-Holland, 1994; pp 3−50. (45) Woodland, A. B.; Koch, M. Earth Planet. Sci. Lett. 2003, 214, 295− 310. (46) Sokol, A. G.; Palyanova, G. A.; Palyanov, Y. N.; Tomilenko, A. A.; Melenevsky, V. N. Geochim. Cosmochim. Acta 2009, 73 (19), 5820− 5834. (47) Lobanov, S. S.; Chen, P.-N.; Chen, X.-J.; Zha, C.-S.; Litasov, K. D.; Mao, H.-K.; Goncharov, A. F. Nature Communications 2013, 4, 1−7. (48) Tomilenko, A. A.; Chepurov, A. I.; Palyanov, Y. N.; Pokhilenko, L. N.; Shebanin, A. P. Geol. Geofiz. 1997, 38, 276−285. (49) Ghodbane, S.; Omnes, F.; Agnes, C. Diamond Relat. Mater. 2010, 19, 273−278. (50) Wang, X.; Ruslinda, A. R.; Ishiyama, Y.; Ishii, Y.; Kawarada, H. Diamond Relat. Mater. 2011, 20, 1319−1324. (51) Zedgenizov, D. A.; Harte, B. Chem. Geol. 2004, 205, 169−175. (52) Howell, D.; Griffin, W. L.; Piazolo, S.; Say, J. M.; Stern, R. A.; Stachel, T.; Nasdala, L.; Rabeau, J. R.; Pearson, N. J.; O’Reilly, S. Y. Am. Mineral. 2013, 98, 66−77. (53) Shatskii, V. S.; Rylov, G. M.; Efimova, E. S.; de Corte, K.; Sobolev, N. V. Geol. Geofiz. 1998, 39, 942−955. (54) Shiryaev, A. A.; Izraeli, E. S.; Hauri, E. H.; Zakharchenko, O. D.; Navon, O. Russ. Geol. Geophys. 2005, 46 (12), 1185−1201. (55) Wiggers de Vries, D. F.; Bulanova, G. P.; De Corte, K.; Pearson, D. G.; Craven, J. A.; Davies, G. R. Geochim. Cosmochim. Acta 2013, 100, 176−199.

The authors declare no competing financial interest.



REFERENCES

(1) Fersman, A.; Goldschmidt, V. Der Diamant; Winter: Heidelberg, Germany, 1911; p 274. (2) Orlov, Yu. L. The Mineralogy of Diamond; John Wiley: New York, 1977; p 235. (3) Sunagawa, I. In Materials Science of the Earth’s Interior; Sunagawa, I., Ed.; Terra Scientific Publishing Company: Tokyo, 1984; pp 303−330. (4) Kvasnitsa, V. N. Fine Diamond; Naukova Dumka: Kiev, SSSR, 1985; p 216 (in Russian). (5) Kaminsky, F. V.; Khachatryan, G. K. Lithos 2004, 77, 255−271. (6) Gurney, J. J.; Hildebrand, P. R.; Carlson, J. A.; Fedortchouk, Ya.; Dyck, D. R. Lithos 2004, 77, 21−38. (7) Lang, A. R. In The Properties of Diamond, Field, J. E., Ed.; Academic Press: London, 1979; pp 425−469. (8) Moore, M.; Lang, A. R. Philos. Mag. 1972, 26, 1313−1326. (9) Welbourn, C. M.; Rooney, M. L. T.; Evans, D. J. F. J. Cryst. Growth 1989, 94, 229−252. (10) Rondeau, B.; Fritsch, E.; Moore, M.; Thomassot, E.; Sirakian, J. F. J. Cryst. Growth 2007, 304, 287−293. (11) Wang, W.; Moses, T. M.; Shigley, J. E. Diamond Relat. Mater. 2003, 12, 330−335. (12) Peats, J.; Stachel, T.; Stern, R. A.; Muehlenbachs, K.; Armstrong, J. Can. Mineral.. 2012, 50, 611−624. (13) Bulanova, G. P. J. Geochem. Explor. 1995, 53, 1−23. (14) Skuzovatov, S. Y.; Zedgenizov, D. A.; Ragozin, A. L.; Shatsky, V. S. Russ. Geol. Geophys. 2012, 53, 1197−1208. (15) Skuzovatov, S. Y.; Zedgenizov, D. A.; Shatsky, V. S.; Ragozin, A. L.; Kuper, K. E. Russ. Geol. Geophys. 2011, 52, 85−96. (16) Zedgenizov, D. A.; Harte, B.; Shatsky, V. S.; Politov, A. A.; Rylov, G. M.; Sobolev, N. V. Contrib. Mineral. Petrol. 2006, 151, 45−57. (17) Bovenkerk, H. P.; Bundy, F. P.; Hall, H. T.; Strong, H. M.; Wentorf, R. H. Nature 1959, 184, 14−18. (18) Giardini, A. A.; Tydings, J. E. Am. Mineral. 1962, 47, 1393−1421. (19) Pal’yanov, Y. N.; Sokol, A. G.; Borzdov, Y. M.; Khokhryakov, A. F.; Sobolev, N. V. Dokl. Akad. Nauk 1998, 363, 230−233. (20) Palyanov, Y. N.; Borzdov, Y. M.; Khokhryakov, A. F.; Kupriyanov, I. N.; Sobolev, N. V. Earth Planet. Sci. Lett. 2006, 250, 269−280. (21) Palyanov, Y.; Borzdov, Y.; Kupriyanov, I.; Gusev, V.; Khokhryakov, A.; Sokol, A. Diamond Relat. Mater. 2001, 10, 2145−2152. (22) Palyanov, Yu. N.; Kupriyanov, I. N.; Sokol, A. G.; Khokhryakov, A. F.; Borzdov, Yu. M. Cryst. Growth Des. 2011, 11, 2599−2605. (23) Pal’yanov, Y. N.; Sokol, A. G.; Borzdov, Y. M.; Khokhryakov, A. F.; Sobolev, N. V. Nature 1999, 400, 417−418. (24) Palyanov, Y. N.; Sokol, A. G. Lithos 2009, 112, 690−700. (25) Shirey, S. B.; Cartigny, P.; Keshav, S.; Nestola, F.; Nimis, P.; Pearon, D. G.; Sobolev, N. V.; Walter, M. J. Rev. Mineral. Geochem. 2013, 75, 355−421. (26) Dobrzhinetskaya, L. F. Gondvana Res. 2012, 21, 207−223. (27) Zhang, Y.; Zang, C.; Ma, H.; Liang, Z.; Zhou, L.; Li, S.; Jia, X. Diamond Relat. Mater. 2008, 17, 209−211. (28) Palyanov, Yu. N.; Borzdov, Yu. M.; Khokhryakov, A. F.; Kupriyanov, I. N.; Sokol, A. G. Cryst. Growth Des. 2010, 10, 3169−3175. (29) Palyanov, Yu. N.; Borzdov, Yu. M.; Kupriyanov, I. N.; Khokhryakov, A. F. Cryst. Growth Des. 2012, 12, 5571−5578. (30) Liu, X.; Jia, X.; Zhang, Z.; Li, Y.; Hu, M.; Zhou, Z.; Ma, H. Cryst. Growth Des. 2011, 11, 3844−3849. (31) Li, Y.; Jia, X.; Hu, M.; Yan, B.; Zhou, Z.; Fang, C.; Zhang, Z.; Ma, H. Int. J. Refract. Met. Hard Mater. 2012, 34, 27−31. (32) Zhang, Z. F.; Jia, X. P.; Sun, S. S.; Liu, X. B.; Li, Y.; Yan, B. M.; Ma, H. A. Int. J. Refract. Met. Hard Mater. 2013, 38, 111−117. (33) Kanda, H.; Setaka, N.; Ohsawa, T.; Fukunaga, O. J. Cryst. Growth 1982, 60, 441−444. (34) Kanda, H.; Setaka, N.; Ohsawa, T.; Fukunaga, O. In Materials Research Society Proceedings; Hofman, C., MacCrone, R. C., Whalley, E., Eds.; Elsevier: Amsterdam, 1984; pp 209−212. 5419

dx.doi.org/10.1021/cg4013476 | Cryst. Growth Des. 2013, 13, 5411−5419