History and Utility of Zeolite Framework-Type ... - ACS Publications

May 2, 2016 - approach, we find that the discovery of exceptional zeolite framework types around the ..... To test whether the dip is a long-term deve...
0 downloads 9 Views 1MB Size
Subscriber access provided by University of Sussex Library

Perspective

History and Utility of Zeolite Framework-Type Discovery from a Data-Science Perspective Nils E.R. Zimmermann, and Maciej Haranczyk Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.6b00272 • Publication Date (Web): 02 May 2016 Downloaded from http://pubs.acs.org on May 6, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Crystal Growth & Design is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

History and Utility of Zeolite Framework-Type Discovery from a Data-Science Perspective Nils E. R. Zimmermann∗ and Maciej Haranczyk Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States E-mail: [email protected] Abstract Mature applications such as fluid catalytic cracking and hydrocracking rely critically on early zeolite structures. With a data-driven approach, we find that the discovery of exceptional zeolite framework types around the new millennium was spurred by exciting new utilization routes. The promising processes have yet not been successfully implemented (“valley of death” effect), mainly because of lacking thermal stability of the crystals. This foreshadows limited deployability of recent zeolite discoveries that were achieved by novel crystal synthesis routes.

Introduction Zeolites are nanoporous, crystalline materials, which are used since many decades in applications ranging from catalysis over gas separation to ion-exchange. 1 Zeolite materials are found in nature (as minerals), and they can be synthesized in the laboratory. The peculiar feature of this class of inorganic materials is that their crystal structure consists of interconnected TO4 tetrahedra, 2 where T stands usually for Si and/or Al. Chemically more diverse materials exist, in which silicon and aluminum are, for example, substituted with P, Fe, and Be. 1

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

However, high-silica materials are most important from an industrial point of view because they typically possess exceptionally high thermal and hydrothermal stability under process conditions. 3 The first scientific report of a zeolite dates back to Cronstedt in 1765, who discovered the mineral stilbite, which he described as “boiling stones”; hence, the word: zeo originates from the Greek word for to boil and lithos is Greek for stone. Already in 1862, St. Claire Deville tried to synthesize zeolites in the laboratory. 4,5 However, the most influential early synthesis milestones from an industrial stance were made by Milton and Breck at Union Carbide in the 1950s. They developed the reactive gel crystallization, which is today perceived as the standard synthesis method for zeolites. The procedure led to the discovery of Al-rich zeolites A and X. Breck, furthermore, claimed the discovery of zeolite Y in 1964. 6 The material is exceptional because zeolite Y is so widely used in fluid catalytic cracking that it represents the most consumed zeolite catalyst. 1 In 1969, ZSM-5, a high-silica zeolite, was synthesized for the first time by Argauer and Landolt. 7 ZSM-5 is outstanding because it is (probably) the zeolite catalyst implemented in the largest number of different processes. 8 Between the late 1970s and 1990s, synthesis of compositionally more exotic structures were targeted. The trend culminated in the spectacular discovery of a synthesis route to chalcogenide zeolite analogs. 9 While the quest for such exceptional structures might seem to have been stimulated by mere scientific curiosity, the main driving force remained the goal of exploiting the new materials for industrial and commercial purposes. 9,10 The success story of SAPO-34, 11 discovered in 1984, underlines this claim. Despite the fact that SAPO-34 has an unconventional composition (silicoaluminophosphate) for being used as an acid-based zeolite catalyst, it led to the world’s first coal-to-olefin plant in 2010. 12 Because of the increasing number of novel materials and because some of those exhibited the same general structure so that they differed in composition only, there appeared the necessity to categorize and group the different zeolite materials. Therefore, the concept of a framework type was introduced. A framework type comprises all materials whose primary

2

ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

building blocks (usually TO4 tetrahedra) are connected in the same way; that is, materials with the same topology. For example, ZSM-5 13 and silicalite 14 can assume very different compositions (Si/Al: 2.7 vs ∞), but both belong to the MFI framework type (Figure 1). The Structure Commission of the International Zeolite Association 15 (IZA), established in 1977, is the official body put in charge by the International Union of Pure and Applied Chemistry (IUPAC) to keep track of discoveries of new framework types. IZA furthermore monitors verifiable discovery and synthesis of new minerals and materials to associate the new compounds with existing framework types. The up-to-date list of known and confirmed frameworks and their related materials is publicly available since 2001 at www.iza-online.org. 16 Each framework type is given a three-letter code based on the name of the first material possessing the framework type in question. For example, the code MFI is derived from the material ZSM-5 because it stands for “Zeolite Socony M obil fi ve”.

MFI

FAU

BEA

MOR

CZP

OSO

RWY

IPC-10

Figure 1: Selected zeolite framework structures.

The usefulness of zeolites can be divided into two categories. On the one hand, there are several applications in which zeolites are actively being used, the most important of which have already been given at the beginning of this Perspective. We will call them “mature technologies” in the following. 17 On the other hand, some zeolites have certain properties that make them promising candidates for other applications. The separation 3

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of enantiomers 5 and photocatalysis 9 might be viewed as the most attractive and, thus, most pursued avenues. Hence, we call them “potential technologies” to underline that the applications seem generally possible; however, there are still unresolved challenges that have prevented wide-spread commercial implementation. Inspired by recent synthesis breakthroughs, 18,19 we pursue here a data-driven approach to uncover trends in zeolite framework-type discovery. We analyze publication statistics and correlate structural and thermodynamic descriptors with individual framework-type discovery years. We find four main phases of zeolite discoveries. Particularly, the third phase carries clear signatures of a targeted quest for structures in potential technologies, and a related trend reversal. The approach yields thus insights into possible driving and correctional forces behind past discoveries. Finally, we argue that a new phase of zeolite discovery is just taking place, and that our approach may help steer those efforts toward practically most useful materials.

Methodology We retrieved 229 crystallographic information files on September 9, 2015, from the IZA website, 16 which provides all known and confirmed frameworks as geometrically optimized purely siliceous structures. The structures were relaxed employing the core-shell force field developed by Sanders, Leslie, and Catlow 20 without symmetry constraints 21 using the General Utility Lattice Program (GULP: version 4.2.3). 22 The procedure can be viewed as the standard protocol when zeolites are being compared in a compositionally “fair” way. 19,21,23 We discarded twelve structures because they were not strictly four-connected (cf., ref 21). The framework discovery year is identified as the earliest report of any material listed by IZA that exhibits the framework type in question. IZA’s requirements for associating a material with a framework type are quite stringent. As a result, the framework discovery year does not necessarily reflect the first-ever scientific report on a material of that framework

4

ACS Paragon Plus Environment

Page 4 of 19

Page 5 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

type. The situation applies in particular to naturally occurring materials. Consider, for example, stilbite: while the mineral was discovered and officially described by Cronstedt in 1765, the official discovery of the underlying STI framework occurred in 1966 by Galli and Gottardi. 24 The discovery year of such frameworks is thus a reflection of the historical efforts to investigate the/a related material in more detail. We believe that the main motivation for more detailed investigations are practical usefulness of a material. Hence, the discovery of such frameworks was likely driven (and the discovery year thus slightly biased) by the exploitability potential of that material. We correlate the discovery year of each framework with two descriptors. The first descriptor, θtet , effectively measures the deviation of angles in a structure motif (here: a given SiO4 unit) away from the perfect tetrahedral angle of αtet = 109.47◦ , for which reason θtet is purely structural. The descriptor was recently introduced by Zimmermann et al. 25 as a local order parameter (OP) for simulation studies of nucleation and polymorph selection and is given by

θtet

4 −(αk − αtet )2 1 X exp = 24 j6=k 2 ∆α2

(

"

#

4 X

−(αm − αtet )2 cos(3 ϕ) exp 2 ∆α2 m6=j,k "

#)

.

(1)

where αk is the angle formed between the bonds of O atoms j and k with their shared Si atom i, ϕ the dihedral angle between bond i–m with the plane spanned by i, j, and k, and ∆α = 12◦ a parameter controlling the reward loss for increasingly nonideal positions. The advantage of this definition of tetrahedrality over existing ones 21,26 is that θtet becomes dimensionless and normalized. Thus, the descriptor can take values between zero and one, where one indicates a perfect tetrahedral coordination environment and zero an extremely distorted surrounding. Using pymatgen 27 for structure analysis we, first, calculate OP values from all individual SiO4 building blocks in each relaxed framework structure; subsequently, we average those to yield a single, representative descriptor per structure (θ¯tet ). At this point, we emphasize that Li et al. 21 established the degree of tetrahedrality in 2013 as one

5

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 19

of the most important factors in rating the feasibility of synthesizing hypothetical zeolite structures. The authors put forth four criteria, which are all based on local interatomic distances (LIDs) as well as variances of and correlations between those. Given that Li et al. 21 introduced the LIDs only recently, we do not exaggerate in noting that the criteria have quickly proven so useful that they are three years later considered a standard. 19,28 The second descriptor that we use to unravel trends in zeolite framework-type discovery is the energy relative to α-quartz:

∆Ei = Ei /NSiO2 ,i − Eqtz /NSiO2 ,qtz

(2)

where Ei denotes the final framework (or lattice) energy from the relaxation of structure i, NSiO2 ,i the number of SiO2 units in the unit cell, and subscript “qtz” indicates α-quartz. The relative lattice energy is a standard thermodynamic descriptor to make predictions about zeolite feasibility. It is typically correlated with the framework density, FD, which gives the number of T-atoms (here: Si) per volume, and which is usually presented in [0.001/Å3 ]. 19,23,29–31 The first time that the correlation between framework energy and density was systematically applied was probably in 1991 by Kramer et al. 29 We will later rationalize why we decouple these two descriptors and focus mainly on the relative framework energy. The two descriptors can be viewed as increasing in complexity. The next level of complexity that reflects relevant information on a material’s feasibility is the consideration of kinetic effects. In this context, Blatov et al. 32 have developed a useful set of criteria based on a model that views zeolite assembly as packings of natural building units and essential rings. Hence, the criteria, once reformulated to yield a more compact rating of kinetic feasibility, may become valuable in monitoring zeolite framework-type discoveries. A possible avenue is to combine the ideas of Blatov et al. 32 with molecular simulation approaches 25 to calculate a typical (effective 33 ) interfacial energy 34 of a given framework structure under some reference synthesis condition (e.g., hydrothermal).

6

ACS Paragon Plus Environment

Page 7 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Results The time-evolution of the framework discoveries per year indicates four different phases (top panel of Figure 2). Because we are interested in long-term trends, we mainly base the analysis on the five-year rolling average, which effectively and sensibly eliminates noise. The first frameworks—ANA, CAN, NAT, SOD—were discovered in 1930 by Taylor 35 and Pauling. 36,37 The most striking feature from the early days is the signature of World War II: between 1934 and 1950 no new zeolite framework were found (or reported on in enough detail). After this first phase, framework discoveries increased steadily until the mid 1960s. At that point, the discovery rate became roughly constant at 1.7/year, with a subtle regressive tendency however. In the early 1980s, the third phase was initiated by a short boom that was followed by another steady period (≈1984 through 1994). During this time, 4.3 structures were discovered on an average per year. In the late 1990s and the early 2000s, two clear peaks in discovery rate are appreciable with an all-time maximum of 14 framework types in 1997. Neglecting some oscillations in the five-year average, Figure 2 suggests that approximately six new framework types have been discovered per year in the fourth phase (late 1990s–today). The discovery process seems to have recently reached another steady state. Our analysis thus quantitatively supports a statement by Payra and Dutta from 2003: 2 “The evolution of materials development in the zeolite field over the last 50 years has followed a path of steady progress, along with steady leaps that introduce new paradigms of synthesis.” Importantly, the current state (since 2008) is unprecedented because the annual discovery rate and the five-year rolling average hardly differ for as much as eight consecutive years now. In contrast to the discovery rate, hardly any long-term trends can be identified with the typical degree in tetrahedrality of newly discovered framework types (thick dark blue line in center panel of Figure 2). Deviations of the descriptor from its all-time mean (hθ¯tet i = 0.976) are small—with two exceptions. First, the average is more sensitive toward noise because of data scarcity at the beginning of discovery phase II. We, therefore, believe that 7

ACS Paragon Plus Environment

Tetrahedral order parameter − θtet

40

20

15 10 5 0

1

0.96

19

19

80

60

II

Page 8 of 19

20

00

19

19

I

20

20

III

IV 5−year average

THO NAT EDI

− 〈 θtet 〉

SOD

− max(θtet) 5−year average (removed CZP, RWY)

VFI

min − (θ t

0.92

GOO

et )

IPC−9

SOS

CDO CZP

0.88

IPC−10

0.84

RWY ∆E 100

100 Energy relative to α−quartz ∆E / [kJ/mol]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Discovery rate [frameworks/yr]

Crystal Growth & Design

75

RWY

0 10 20 Framework density 3 FD / [0.001 T/Å ]

50

25

0

EDI NAT THO

max(∆E)

〈∆E 〉 19

20

O

OS

LOV

IPC−10 IPC−9

min(∆E ) 19

40

19

19

60

T

JS

CZP WEI AFY

80

20

00

20

20

Year

Figure 2: Historical evolution of descriptors related to zeolite framework-type discovery: annual discovery rate (top), average tetrahedrality of SiO4 building blocks in a given framework (θ¯tet : center), and lattice energy relative to α-quartz (∆E: bottom). Gray histograms (top) and circles (center and bottom) represent raw data. To eliminate noise and uncover long-term trends we additionally display five-year rolling averages of the descriptors (solid dark blue lines). Because typical funding cycles are around five years, the choice of a fiveinstead of a three- or a seven-year rolling average seems appropriate. We furthermore indicate all-time averages (h. . .i: dashed green lines) as well as the all-time minimum and maximum as envelopes for the structural and thermodynamic descriptor (solid light blue lines). The inset of the bottom panel shows the energy-density landscape of all known zeolite frameworks. 19,23,29–31 Note that there is no distinct correlation between (∆E, FD) and θ¯tet .

the oscillations between 1950 and 1963 do not indicate statistically meaningful trends. By contrast, the second evident feature—a dip toward more distorted tetrahedra between 1999 and 2008—does seem systematic. To test whether the dip is a long-term development or caused by a few exceptions we consider the extrema envelopes of that descriptor (min(θ¯tet ) 8

ACS Paragon Plus Environment

Page 9 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

and max(θ¯tet )). The all-time maximum in tetrahedrality does not change significantly over time because highly ideal structures (NAT, THO, EDI) were already among the earliest frameworks discovered. However, the time-evolution of the all-time minimum suggests that there was a concrete quest for new zeolite materials that are structurally extraordinary. The time between two subsequent record minima continuously decreased up to the discovery of RWY in 2002. When we remove the two possible outsiders RWY and CZP from the rolling average, the decreasing trend and the subsequent reversal remain (purple line). This further supports the hypothesis of large concerted efforts to find exceptional frameworks, followed by a change in strategy away from those exotic structures. Similar observations can be made for the relative framework energy (bottom panel of Figure 2). The five-year average of ∆E deviates however more evidently from the global mean h∆Ei compared to the evolution of θ¯tet . The deviations might be signatures of medium-term trends. Because the descriptor is usually correlated with the framework density and because that correlation is not very strong, we believe that those weak intermediate trends in ∆E should be viewed with caution. The striking feature of a sharp increase in 1999 and a temporal plateau until 2008 seem yet to confirm the peculiar trend unraveled by the structural descriptor θ¯tet . Also, a general trend toward high-energy structures since the late 1970s cannot be denied.

Discussion Discovery phase III is frequently viewed as mainly driven by achieving compositionally exotic materials. And, the discovery of the chalcogenide zeolite analogs 9 is clear evidence for these efforts. Those materials established framework type RWY, and they are in more than one way exceptional: compositionally as well as structurally and thermodynamically as our descriptor analysis underlines. The technological interest behind the efforts are rooted in the quest for semiconducting porous materials that would have photocatalytic activity. 9 However, there

9

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

is another group of exceptional materials en route to RWY whose discovery pushed the structural and thermodynamic limit of feasible zeolites in a similar manner: chiral structures (e.g., CZP, GOO, OSO). Chirality—or, handedness—in zeolites was believed to be exploitable for the separation of enantiomeric adsorbates since the influential paper by Davis et al. in 1992. 5 The biggest challenge in this context is the targeted synthesis of a single enantiomer of the chiral structure in question. 38,39 Organic structure directing agents (OSDA) are typically used in the synthesis. The OSDAs have to be removed to enable enantiomeric separation in the first place. Therefore, thermal stability is the second big hurdle that has prevented chiral zeolites from establishing a new mature separation technology. CZP 40 and OSO 41 are good examples of the second hurdle because both frameworks possess very limited thermal stability. 42,43 Importantly, both structures exhibit low tetrahedrality (0.896, 0.951) and high relative framework energies (40, 54). Similarly, semiconducting zeolites for photocatalytic purposes could not be established as a new mature technology since the discovery of RWY 13 years ago. 9 Although new materials of (probably) new framework types are still being found and assessed as in-principle promising (photo-)electrocatalyst, the main problem, apart from the synthesis, is also long-term thermal stability under process-like conditions. 18 Insufficient thermal stability correlates with the simple descriptors (low θ¯tet and high ∆E). For those framework types employed in the most mature and important technologies, 8,17 we see that the descriptor values are close to the averages from all 217 structures (Table 1). Hence, the descriptors could become valuable tools in rating not only the feasibility of hypothetical frameworks, but also the likelihood to exploit existing structures in mature future technologies. It is tempting to interpret the trend reversal around 2002 as an active response of the synthesis community based on our discussion. A reasonable hypothesis seems to be that the lack of implementing exotic structures as thermally stable materials for new mature technologies

10

ACS Paragon Plus Environment

Page 10 of 19

Page 11 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Table 1: Structural and thermodynamic descriptor for industrially relevant zeolite frameworks and averages from all frameworks. Framework Tetrahedrality Relative Energy type θ¯tet ∆E/[kJ/mol] MFI 0.979 10 FAU 0.979 20 BEA 0.976 14 MOR 0.971 12 CHA 0.977 16 Average 0.976 17

has prevented further exploration of that region of material space. However, synthesis on demand is still a long shot and the broader goal of somewhat rational synthesis design is just becoming available. 39 The recent discovery of new framework structures IPC-9 and IPC-10 by Mazur et al. 19 showcases the more likely reason behind the trend reversal: new synthesis paradigms that are being discovered. 2 Specifically, Mazur et al. 19 used an elaborate multistep synthesis route called ADOR (assembly-disassembly-organization-reassembly). The two structures thus discovered are thermodynamically not too exceptional when all framework types are compared (orange symbols in bottom panel of Figure 2). But IPC-10 is extraordinary from a structure point of view (center panel). In light of the exciting discoveries from 2015, 18,19 we wish to comment on Mazur et al. 19 who “envisage that (their) findings may lead to a step change in the number and types of zeolites available for future applications.” Our Perspective comes at a time when another zeolite framework-type discovery phase might just have started: yes. And, we see multiple indications for a new (type of) discovery phase: exceptionally stable discovery rates as well as forerunners of a new wave of structurally exotic materials. However, Schoonover and Cohn 44 noted in 2000 already that there is a large discrepancy between the discovery rate of new zeolite materials and the rate of commercialization. The effect is largely known as the “valley of death”. 45 Underlying the “valley of death” is a view on technology transfer from scientific proof-of-concept to concrete commercial deployment (Figure 3). The transfer

11

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 19

is pictured as a cascade of transitions from one milestone (valley) to the next milestone. The transitions are impeded by barriers (hills). Frequently, one barrier is so high that the technology transfer is practically unachievable; hence, the technology “dies” in the “valley” located ahead of the (first) insurmountable “transfer hill.”

Proof-of-concept

Valle of dea

e

th

n Sy

ce

sis

Pe

m

r rfo

an

(Commercial eployment)

ce an

ic

om on ty Ec bili via

r

du

En

Technology transfer progress

Figure 3: Technology transfer is typically viewed as a cascade of barriers hindering the successful move from proof-of-concept to commercial deployment. In the vast majority of cases, one of the barriers cannot be overcome (here: endurance or thermal stability under process conditions), giving the preceding well its incisive name: “valley of death.”

Schoonover and Cohn 44 attributed three basic hurdles to commercial exploitability: new material discovery, satisfactory performance, and economic utility. The economical dimension is often insurmountable; for example, it has likely suppressed the substitution of existing acidic zeolite catalysts (ZSM-5) with novel zeolitic materials (MWW 46 types). However, before that stage is reached, central performance tests have to be passed. Part of the satisfactory performance is long-term stability under process conditions. Given that rational synthesis design is becoming more and more available, 39 we believe that it is critical to actively intervene into the “next zeolite revolution” right from the onset. 47 In particular, we wish to help steer synthesis efforts into the direction that is most likely most useful: toward satisfactory performance under process conditions. Based on our insights, the goal should be to realize materials with an optimal degree in tetrahedrality (θ¯tet ) and thermodynamic stability (∆E) because the ultimate goal is not “to adjust external conditions in such a way

12

ACS Paragon Plus Environment

Page 13 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

that” new zeolite frameworks “can be realized or not.” 48,49 The ultimate goal is to deploy the new structures under typically harsher conditions than during synthesis in order to be exploited in a practically meaningful way.

Conclusions Current synthesis breakthroughs 18,19 have triggered the discovery of new exotic zeolite structures, marking the scientifically valuable exploration of a large new region of materials space. Viewed from a pragmatic stance that takes into account past trends in this area, following question appears however. Should we actively strive again for highly distorted, high-energy zeolite materials if the ultimate goal was, is, and always will be materials utility, given that the past has revealed this direction as ineffective for triggering a step change in materials truly deployable in new types of applications?

Acknowledgement We thank the anonymous reviewer of this Perspective for suggesting discussions of technological substitution barriers and SAPO-34. This work was supported by the U.S. Department of Energy, Office of Basic Energy Sciences, Division of Chemical Sciences, Geosciences and Biosciences, under award DE-FG02-12ER16362. Lawrence Berkeley National Laboratory is funded by the U.S. Department of Energy under award DE-AC02-05CH11231.

Supporting Information Available Movie summarizing our Perspective. This material is available free of charge via the Internet at http://pubs.acs.org/.

13

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) Flanigen, E. M.; Broach, R. W.; Wilson, S. T. Zeolites in Industrial Separation and Catalysis; Wiley-VCH Verlag: Weinheim, 2010; Chapter Introduction, pp 1–26. (2) Payra, P.; Dutta, P. K. Handbook of Zeolite Science and Technology; Marcel Dekker, Inc.: New York, NY, USA, 2003; Chapter Zeolites: a primer, pp 1–24. (3) Morris, R. E.; Čejka, J. Exploiting chemically selective weakness in solids as a route to new porous materials. Nat. Chem. 2015, 7, 381–388. (4) de St. Claire Deville, H. Compt. Rend. 1862, 54, 324. (5) Davis, M. E.; Lobo, R. F. Zeolite and molecular sieve synthesis. Chem. Mater. 1992, 4, 756–768. (6) Breck, D. W. Crystalline zeolite Y. U.S. Patent 3130007, 1964. (7) Argauer, R. J.; Landolt, G. R. Crystalline zeolite ZSM-5 and method of preparing the same. U.S. Patent 3702886 A, 1972. (8) Liu, Z.; Wang, Y.; Xie, Z. Thoughts on the future development of zeolitic catalysts from an industrial point of view. Chinese J. Cat. 2012, 33, 22–38. (9) Zheng, N.; Bu, X.; Wang, B.; Feng, P. Microporous and photoluminescent chalcogenide zeolite analogs. Science 2002, 298, 2366–2369. (10) Corma, A.; Díaz-Cabañas, M. J.; Martínez-Triguero, J.; Rey, F.; Rius, J. A large-cavity zeolite with wide pore windows and potential as an oil refining catalyst. Nature 2002, 418, 514–517. (11) Lok, B. M.; Messina, C. A.; Patton, R. L.; Gajek, R. T.; Cannan, T. R.; Flanigen, E. M. Silicoaluminophosphate molecular sieves: another new class of microporous crystalline inorganic solids. J. Am. Chem. Soc. 1984, 106, 6092–6093. 14

ACS Paragon Plus Environment

Page 14 of 19

Page 15 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(12) Tian, P.; Wei, Y.; Ye, M.; Liu, Z. Methanol to olefins (MTO): from fundamentals to commercialization. ACS Catal. 2015, 5, 1922–1938. (13) Kokotailo, G. T.; Lawton, S. L.; Olson, W. M., D. H. Meier Structure of synthetic zeolite ZSM-5. Nature 1978, 272, 437–438. (14) Flanigen, E. M.; Bennett, J. M.; Grose, R. W.; Cohen, J. P.; Patton, R. L.; Kirchner, R. M.; Smith, J. V. Silicalite, a new hydrophobic crystalline silica molecular sieve. Nature 1978, 271, 512–516. (15) Baerlocher, C.; McCusker, L. B.; Olsen, D. H. Atlas of Zeolite Framework Types, 6th ed.; Elsevier: Amsterdam, The Netherlands, 2007. (16) Baerlocher, C.; McCusker, L. B. Database of Zeolite Structures. http://www. iza-structure.org/databases/. (17) Abate, S.; Barbera, K.; Centi, G.; Lanzafame, P.; Perathoner, S. Disruptive catalysis by zeolites. Catal. Sci. Technol. 2016, (18) Lin, J.; Dong, Y.; Zhang, Q.; Hu, D.; Li, N.; Wang, L.; Liu, Y.; Wu, T. Interrupted chalcogenide-based zeolite-analogue semiconductor: atomically precise doping for tunable electro-/photoelectrochemical properties. Angew. Chem., Int. Ed. 2015, 54, 5103– 5107. (19) Mazur, M.; Wheatley, P. S.; Navarro, M.; Roth, W. J.; Položij, M.; Mayoral, A.; Eliášová, P.; Nachtigall, P.; Čejka, J.; Morris, R. E. Synthesis of ’unfeasible’ zeolites. Nat. Chem. 2016, 8, 58–62. (20) Sanders, M. J.; Leslie, M.; Catlow, C. R. A. Interatomic potentials for SiO2 . J. Chem. Soc., Chem. Commun. 1984, 1271–1273. (21) Li, Y.; Yu, J.; Xu, R. Criteria for zeolite frameworks realizable for target synthesis. Angew. Chem., Int. Ed. 2013, 52, 1673–1677. 15

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(22) Gale, J. D. GULP: Capabilities and prospects. Z. Kristallogr. 2005, 220, 552–554. (23) Majda, D.; Paz, F. A. A.; Friedrichs, O. D.; Foster, M. D.; Simperler, A.; Bell, R. G.; Klinowski, J. Hypothetical zeolitic frameworks: in search of potential heterogeneous catalysts. J. Phys. Chem. C 2008, 112, 1040–1047. (24) Galli, E.; Gottardi, G. Miner. Petrogr. Acta 1966, 12, 1–10. (25) Zimmermann, N. E. R.; Vorselaars, B.; Quigley, D.; Peters, B. Nucleation of NaCl from aqueous solution: critical sizes, ion-attachment kinetics, and rates. J. Am. Chem. Soc. 2015, 137, 13352–13361. (26) Zwijnenburg, M. A.; Simperler, A.; Wells, S. A.; Bell, R. G. Tetrahedral distortion and energetic packing penalty in “zeolite” frameworks: Linked phenomena? J. Phys. Chem. B 2005, 109, 14783–14785. (27) Ong, S. P.; Richards, W. D.; Jain, A.; Hautier, G.; Kocher, M.; Cholia, S.; Gunter, D.; Chevrier, V. L.; Persson, K. A.; Ceder, G. Python Materials Genomics (pymatgen): A robust, open-source python library for materials analysis. Comput. Mater. Sci. 2013, 68, 314–319. (28) Liu, X.; Valero, S.; Argente, E.; Botti, V.; Sastre, G. The importance of T—T—T angles in the feasibility of zeolites. Z. Kristallogr. - Cryst. Mater. 2015, 230, 291–299. (29) Kramer, G. J.; de Man, A. J. M.; van Santen, R. A. Zeolites versus aluminosilicate clusters: the validity of a local description. J. Am. Chem. Soc. 1991, 113, 6435–6441. (30) Henson, N. J.; Cheetham, A. K.; Gale, J. D. Theoretical calculations on silica frameworks and their correlation with experiment. Chem. Mater. 1994, 6, 1647–1650. (31) Pophale, R.; Cheeseman, P. A.; Deem, M. W. A database of new zeolite-like materials. Phys. Chem. Chem. Phys. 2011, 13, 12407–12412.

16

ACS Paragon Plus Environment

Page 16 of 19

Page 17 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(32) Blatov, V. A.; Ilyushin, G. D.; Proserpio, D. M. The zeolite conundrum: why are there so many hypothetical zeolites and so few observed? a possible answer from the zeolitetype frameworks perceived as packings of tiles. Chem. Mater. 2013, 25, 412–424. (33) Agarwal, V.; Peters, B. Advances in Chemical Physics: Volume 155 ; John Wiley & Sons, Inc., 2014; pp 97–160. (34) Barrer, R. M. Zeolites and their synthesis. Zeolites 1981, 1, 130–140. (35) Taylor, W. H. The structure of analcite (NaAlSi2O6 H2O). Z. Kristallogr. 1930, 74, 1–19. (36) Pauling, L. The structure of some sodium and calcium aluminosilicates. Proc. Natl. Acad. Sci. USA 1930, 16, 453–459. (37) Pauling, L. The structure of sodalite and helvite. Z. Kristallogr. 1930, 74, 213–225. (38) Davis, M. Ordered porous materials for emerging applications. Nature 2002, 417, 813– 821. (39) Moliner, M.; Rey, F.; Corma, A. Towards the rational design of efficient organic structure-directing agents for zeolite synthesis. Angew. Chem., Int. Ed. 2013, 52, 13880–13889. (40) Rajić, N.; Logar, N. Z.; Kaučič, V. A novel open framework zincophosphate: synthesis and characterization. Zeolites 1995, 15, 672–678. (41) Cheetham, A. K.; Fjellvåg, H.; Gier, T. E.; Kongshaug, K. O.; Lillerud, K. P.; Stucky, G. D. In Stud. Surf. Sci. Catal.; Galarneau, A., Fajula, F., Di Renzo, F., Vedrine, J., Eds.; Elsevier, 2001; Vol. 135; p 158. (42) Harrison, W. T. A.; Gier, T. E.; Stucky, G. D.; Broach, R. W.; Bedard, R. A. NaZnPO4 ·H2 O, an open-framework sodium zincophosphate with a new chiral tetrahedral framework topology. Chem. Mater. 1996, 8, 145–151. 17

ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(43) Strohmaier, K. G.; Vaughan, D. E. W. Structure of the first silicate molecular sieve with 18-ring pore openings, ECR-34. J. Am. Chem. Soc. 2003, 125, 16035–16039. (44) Schoonover, M. W.; Cohn, M. J. New materials discovery for industrial applications. Top. Catal. 2000, 13, 367–372. (45) Branscomb, L. M.; Auerswald, P. E. Taking Technical Risks: How Innovators, Managers, and Investors Manage Risk in High-Tech Innovations; MIT Press: Cambridge, MA, USA, 2001. (46) Puppe, L.; Weisser, J. Crystalline aluminosilicate PSH-3 and its process of preparation. U.S. Patent 4439409 A, 1984. (47) Kusiak, A. Put innovation science at the heart of discovery. Nature 2016, 530, 255. (48) Kuhnert-Brandstätter, M. Polymorphie bei Arzneistoffen. Pharma. Z. 1975, 4, 131– 137. (49) Bernstein, J. Polymorphism — a perspective. Cryst. Growth Des. 2011, 11, 632–650.

18

ACS Paragon Plus Environment

Page 18 of 19

Page 19 of 19

For Table of Contents Use Only Title: History and Utility of Zeolite Framework-Type Discovery from a Data-Science Perspective Authors: Nils E. R. Zimmermann Maciej Haranczyk

Valley of death

Tetrahedrality

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Year

Synopsis We caution high hopes in an on-going zeolite revolution because exotic crystal structures for promising new applications have already been discovered in the past at the cost of very limited thermal stability under process conditions (“valley of death”).

19

ACS Paragon Plus Environment