Este sitio web utiliza cookies. Si continúas navegando estás dando su consentimiento para la aceptación de las mencionadas cookies y la aceptación de nuestra política de cookies.
unvidente.com unvidente.com

unvidente.com
unvidente.com
unvidente.com
  • 0 voto(s) - 0 Media
  • 1
  • 2
  • 3
  • 4
  • 5

Towards understanding the improved stability of palladium supported on TS-1 for catal

Sin conexión
Towards understanding the improved stability of palladium supported on TS-1 for catalytic combustion

A novel Pd supported on TS-1 combustion catalyst was synthesized and tested in methane combustion under very lean and under highly humid conditions (<1%). A notable increase

in hydrothermal stability was observed over 1900 h time-on-stream experiments, where an almost constant, steady state activity obtaining 90% methane conversion was achieved

below 500 °C. Surface oxygen mobility and coverage plays a major role in the activity and stability of the lean methane combustion in the presence of large excess of water

vapour. We identified water adsorption and in turn the hydrophobicity of the catalyst support as the major factor influencing the long term stability of combustion

7% palladium on carbon. While Pd/Al2O3 catalyst

shows a higher turn-over frequency than that of Pd/TS-1 catalyst, the situation reversed after ca. 1900 h on stream. Two linear regions, with different activation energies in

the Arrhenius plot for the equilibrium Pd/TS-1 catalyst, were observed. The conclusions were supported by catalyst characterization using H2-chemisorption, TPD, XPS analyses as

well as N2-adsorption–desorption, XRD, SEM, TEM. The hydrophobicity and competitive adsorption of water with oxygen is suggested to influence oxygen surface coverage and in

turn the apparent activation energy for the oxidation reaction.

The selective hydrogenation of a range of substrates is a key technology in both the bulk and fine chemicals industries [1]. In both contexts, selectivity to the desired

product is usually a key attribute: loss of reagent to the formation of undesired products is economically undesirable and can lead to challenges in separation downstream. This

means that there is a pressing need for more selective catalysts and processes for a range of selective hydrogenation reactions. One way to meet this need is the design and

realization of catalytic materials with improved properties. The majority of commercial [url=http://www.kailicatalyst.com/palladium-series-catalyst/heterogeneous-catalyst-of-

palladium/5-palladium-on-carbon.html]5% palladium on carbon [/url]are made using a small number of synthesis methods (impregnation, precipitation, solid-state methods, etc.).

There is good reason for this: they are reliable, economic, and can be performed at the necessary scale for commercial use. However, they are not always able to produce

materials that are truly optimized.

Making an optimized catalyst requires control over the synthesis of the active site, as well as attachment of the active site to the catalyst support (which is typically

needed for mechanical properties as well as to disperse the active sites). For the former, the use of nanoparticles synthesized in solution is an attractive proposition. They

can be produced ex situ from the catalyst support by controlling the key properties such as particle size [2], shape [3], and the nature of the exposed surfaces [4] and can

contain more than one metal with controlled location (such as a core–shell structure) [5]. Attaching these particles to supports is a complex process. Although in some cases

the presence of stabilizers has been shown to be beneficial [6], often the stabilizers need to be removed for optimal performance. Ligand removal often changes the nature of the

nanoparticle, for example through a loss of size control [7], rendering them poorly performing. Ligand removal has been addressed in a few selected cases, for example in a

catalyst made with polymer-stabilized nanoparticles [8], but significant progress is still needed to find a general method that would allow manufacturing at scale to take place.

Synthesis of nanoparticles by aggregation of metal atoms or ions in the gas phase is a promising technology [9] that addresses many of these issues. In a typical

configuration, atoms are generated from a metal source and these are condensed to form clusters. Typically, some of the particles formed are charged, which allows them to be

manipulated using applied voltages, mass-selected if desired, and finally guided onto the support. The technique can offer particle-size control from less than 2 nm to over 10

nm [10] and also some control over the interaction between the nanoparticle and the support: the accelerating voltage can be used to control the impact of the particle into the

support [11–13]. We [14] and others [15] have, in this way, made bimetallic clusters from a number of metals. Yang et al. [16] have demonstrated the selective deposition of

silver clusters onto the top face of silicon pillars. A combination of these different features should allow the design of catalysts with a high degree of control.

In this work, we use gas-phase cluster deposition as a method to deposit size-controlled palladium series

catalyst
onto two typical commercial powder support materials. We employ the selective partial hydrogenation of 1-pentyne (Scheme 1) as a model reaction for the selective

hydrogenation of alkynes relevant to both the bulk [17] and fine [18,19] chemicals industries. We have previously reported the good performance of a palladium catalyst prepared

by gas-phase cluster deposition onto a flat graphite tape as a catalyst for the selective hydrogenation of 1-pentyne [20], and we have also observed changes in the atomic

structure of size-selected palladium nanoparticles during this reaction [21]. Most recently, we have reported the performance of PdM bimetallic cluster catalysts in alkyne

hydrogenation [14]. In this paper, we describe the performance of catalysts prepared by gas-phase nanoparticle synthesis in selective alkyne hydrogenation and offer some

perspective on the nature of the reactive sites.

Figure 1. Representative bright-field aberration-corrected STEM images of the catalysts prepared by gas-phase cluster deposition: (A)–(B) Pd/α-Al2O3; ©–(D) Pd/TiO2.

Examples of palladium particles are indicated by red arrows, alpha alumina particles with yellow arrows, and titania particles with blue arrows.

Palladium was deposited on two conventional support powders (alpha alumina and titania) to make representative catalysts for the vapor-phase selective hydrogenation of 1-

pentyne to 1-pentene. Catalysts were prepared by four methods: gas-phase cluster beam deposition, incipient wetness impregnation, deposition-precipitation, and ion-exchange

methods. Details of the methods used are presented in the Supporting Information. Table 1 compares the properties of the catalysts. The palladium content of the materials is low

at 0.1wt%. This was driven by the experimental configuration for gas-phase cluster deposition. However, the efficient use of scarce precious metal resources is a key

consideration, and synthetic methods for making good catalysts at these low loadings are valuable. Figure 1 shows representative TEM images of the catalysts synthesized by gas-

phase cluster deposition, whereas images of the other catalysts are presented in Fig. S2. Table 1 lists the particle-size ranges for the catalysts. It was difficult to determine

precise distributions of the nanoparticles due to clustering in some systems and low loading in the others.

In gas-phase cluster deposition on both supports, nanoparticles are observed only close to the support surface, where they often form agglomerates. In the case of titania,

the support is present as a loose agglomerate of 20–30 nm particles, and the palladium particles are deposited on the surface of these agglomerates. The alpha alumina is

present as much larger particles (20–40 μm), and here the heterogeneous

catalyst of palladium
are deposited on the alumina particle surface with little transport of the nanoparticles into the interior of the alumina. Although deposition on the

external surface is a general feature of gas-phase cluster deposition processes, neither the alpha alumina nor the titania used in this work is significantly porous, so the

materials are all expected to be surface enriched in palladium. Clearly, this would not be the case for a more porous support, such as a typical gamma alumina.

Catalyst testing

The catalysts' performance in the selective hydrogenation of 1-pentyne (Scheme 1) were tested in a quartz microreactor using the as-prepared powders. 1-pentyne vapor and

a hydrogen–helium mixture were flowed through a catalyst bed while the temperature was increased from ambient to 250°C. Full details of the catalytic testing methodology are

presented in the Supporting Information. Figure 2 shows the performance of the eight catalysts when tested at equivalent palladium content and bed depth. None of the catalysts

showed a significant amount of activity at low temperature (<50°C). As the temperature increases above this temperature, the 1-pentyne conversion increases. The most active of

the catalysts studied were Pd/α-Al2O3 prepared by impregnation and by deposition-precipitation. The gas-phase cluster deposition materials were the least active, but also the

most selective, with combined selectivity to 2-pentenes and pentane of less than 10% across the temperature range studied.

Figure 2. Catalyst testing in 1-pentyne hydrogenation. The 1-pentyne conversion is shown in blue, with selectivity to 1-pentene (red), 2-pentenes (green, solid line), and

pentane (green, solid line) also shown. The catalysts are (a) Pd/TiO2 GCD; (b) Pd/α-Al2O3 GCD; © Pd/TiO2 impregnation; (d) Pd/α-Al2O3 impregnation; (e) Pd/TiO2 deposition-

precipitation; (f) Pd/α-Al2O3 deposition-precipitation; (g) Pd/TiO2 ion exchange; and (h) Pd/α-Al2O3 ion exchange.

Given the difference in activity between the GCD and reference catalysts, it was of interest to compare their performance at close to iso-conversion. This was achieved by

varying the catalyst mass at constant flow rates of hydrogen and 1-pentyne. Details of the procedure are presented in the Supporting Information. Table 2 shows the selectivity

of each catalyst when the temperature was at a point where 80% conversion was achieved (T80). Under these conditions, the selectivity of the catalysts is much closer, although

the GDC catalysts are still among the best for each support studied. The most selective catalysts are Pd/α-Al2O3 prepared by impregnation, deposition-precipitation, and gas-

phase cluster deposition with over 90% selectivity to 1-pentene. The Pd/TiO2 catalysts are generally less selective. Intriguingly, the two catalysts prepared by the ion-exchange

method have very similar performances.It is clearly of interest to understand the origin of the performance of the eight catalysts studied. The materials present a range of

metal–support interaction types, and these can be used to understand how the nature of the active site affects catalytic performance. For the materials prepared by gas-phase

cluster deposition, there is no contact between Pd2+ ions and the support, whereas for ion-exchange materials, the interaction is governed by the adsorption of Pd2+ ions onto

reactive sites on the support, such as Al-O? or Ti-O?, by the replacement of two H+ ions with one Pd2+ ion. The isoelectric points of alpha alumina and titania are reported to

be pH 9.3 [22] and pH 5.4 [23], respectively. The metal precursor used in this study, [url=http://www.kailicatalyst.com/palladium-series-catalyst/homogeneous-catalyst-of-

palldium/]homogeneous catalyst of palldium [/url]nitrate, is acidic, which makes the impregnating solution acidic. However, even at lower pH, some negatively charged surface

sites will exist [24]; clearly, the number and distribution will be affected by the nature of the palladium precursor solution and the support material. At ion-exchange sites,

palladium will be transformed during subsequent thermal treatments (in this work, drying at 100°C and hydrogen reduction at 250°C). In the final catalyst, they will behave

differently from the main nanoparticulate palladium phase and invariably lead to some loss of selectivity under reaction conditions. If these sites were highly active, as might

be anticipated for a very well-dispersed phase, they could influence selectivity disproportionately.
Sin conexión
Why not settling on games that is fun and at the same time your earning. Well itll make suspense because of the game as well but dude just try it and it gave me hope while pandemic is real rn. Gambling Mistakes That Can Ruin Your Finances
Salto de foro:

Usuarios navegando en este tema: 1 invitado(s)
unvidente.comTarot 806
806 533 561
unvidente.comTarot Visa y PayPal
919 991 039
unvidente.comUSA, Canadá, Pto. Rico
1-305-507-8029
unvidente.comTarot Argentina
+54 (11) 52198820
unvidente.comTarot México
+52 (55) 85266010
unvidente.comResto del Mundo
+34 919 991 039