• No results found

University of Groningen Self-replicators from dynamic molecular networks: selection, competition and subsystem coupling Komáromy, Dávid

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Self-replicators from dynamic molecular networks: selection, competition and subsystem coupling Komáromy, Dávid"

Copied!
43
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Self-replicators from dynamic molecular networks: selection, competition and subsystem

coupling

Komáromy, Dávid

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Komáromy, D. (2019). Self-replicators from dynamic molecular networks: selection, competition and

subsystem coupling. University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the

author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately

and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the

number of authors shown on this cover page is limited to 10 maximum.

(2)

Chapter 2

Self-Assembly Can Direct Dynamic Covalent

Bond Formation toward Diversity or

Specificity

With the advent of reversible covalent chemistry the study of the interplay between

covalent bond formation and noncovalent interactions has become increasingly relevant.

Here we report that the interplay between reversible disulfide chemistry and

self-assembly can give rise either to molecular diversity, i.e., the emergence of a

unprecedentedly large range of macrocycles or to molecular specificity, i.e., the

autocatalytic emergence of a single species. The two phenomena are the result of two

different modes of self-assembly, demonstrating that control over self-assembly

pathways can enable control over covalent bond formation.

This chapter has been published:

D. Komáromy, M. C. A. Stuart, G. Monréal Santiago, M. Tezcan, V. V. Kranikov, S. Otto

(3)

44

2.1. Introduction

Biological systems function by virtue of a complex and concurrent interplay of covalent bond

formation and non-covalent assembly processes. For example, the non-covalent assembly of protein

complexes affects the ability of the protein to catalyze covalent chemical reactions.

[1]

On the other hand,

covalent histone modifications control the non-covalent binding of a DNA strand within the chromatin

complex.

[2]

In chemistry, traditionally the processes of covalent and non-covalent bond formation occur

sequentially: for example, first the potential host and guest molecules are synthesized by covalent

chemical means and subsequently, host-guest binding or self-assembly is investigated. In many cases

such separation of covalent and non-covalent processes cannot be avoided as the conditions required

for organic synthesis are often incompatible with those required for non-covalent interactions. However,

with the advent of reversible covalent chemistry,

[3–5]

the development of systems featuring concurrent

covalent and non-covalent chemistries has become possible. Dynamic covalent systems constructed

from different building blocks tend to lead to diverse mixtures of products which continuously exchange

building blocks via reversible covalent bond formation (dynamic combinatorial libraries, DCLs).

[6–8]

Non-covalent interactions

[9]

can then be utilized to channel the building blocks into specific DCL members that

optimally engage in molecular recognition. This effect has been exploited for the dynamic combinatorial

discovery of synthetic receptors

[10–17]

and ligands for biomolecules

[18,19]

by exposing the system to a

corresponding template.

The theory of template-induced amplification of specific DCL members is well-established.

[20–25]

More

recently, also molecular recognition processes that take place in the absence of added templates,

[26]

occurring between library members have been explored, leading to interlocked structures,

[27–31]

self-

replicating molecules

[32–38]

and self- assembling materials

[39–41]

. In these systems, non-covalent

interactions within (for interlocked structures) or between (for self-replicating and self-assembling

materials) specific library members shift the equilibrium towards molecules that engage most efficiently

in non-covalent interactions. Such behavior is relevant from the perspectives of the origin-of-life

[42–44]

and the de-novo synthesis of life,

[45–49]

as it leads to the spontaneous and often autocatalytic emergence

of specific molecules from complex mixtures, where these molecules had acquired information and are

able to pass this information on to the next generation during self-replication. Self-assembly phenomena

are also intimately linked with materials science, as supramolecular objects based on molecules

containing dynamic covalent bonds undergoing spontaneous assembly can be regarded as

self-synthesizing.

34,[50–52]

Despite the considerable interest in dynamic combinatorial self-assembly, the

number of such systems is limited, particularly when it comes to amphiphile self-assembly.

[39,50,53–55]

Moreover, in contrast to the well-established theory of template-induced amplification,

[20–25]

the

(4)

45

selection remains underdeveloped. We now report how two different self-assembly pathways induce

dramatically different responses in the behavior of libraries made from the same building block. One

pathway leads to a diverse set of unprecedentedly large macrocycles, while a second pathway leads to

the autocatalytic formation of one specific macrocycle.

Scheme 2. 1. Dynamic Combinatorial Chemistry of Building Block 1.

2.2. Results and Discussion

In order to explore self-assembly in dynamic combinatorial chemistry, we designed amphiphilic

building block 1 (Scheme 2. 1), containing a short polar oligo(ethylene oxide) chain connected to a

nonpolar aromatic ring functionalized with two thiol groups for reversible covalent disulfide chemistry.

Under slightly basic conditions dithiols are (partially) deprotonated to give thiolates which are oxidized

slowly by atmospheric oxygen (or faster using sodium perborate) to disulfides. Thiolates also react with

disulfides in a reversible manner, which enables exchange, and thus, dynamic covalent chemistry

between the disulfides.

The oxidation of building block 1 in a 9:1 mixture of aqueous borate buffer (pH= 8.2) and

dimethylformamide (DMF) in the absence of mechanical agitation yielded a DCL which consisted mainly

(5)

46

of cyclic trimers and tetramers, but featured a considerable amount (30 mol%) of larger macrocyclic

species (LMCs) as well, from cyclic 7

mer

to cyclic 44

mer

(Figure 2. 1A and Figure S2. 24).

Figure 2. 1.UPLC analyses of DCLs made from 6.0 mM building block 1 in a 9:1 mixture of aqueous borate buffer (50 mM, pH 8.2) and DMF (A) without agitation and (B) stirred at 1200 rpm.

The identity of the observed species (i.e. cyclic as opposed to linear) was confirmed by LC-MS

analysis, as shown in Section 2.4.6. Although we could not exclude that some of the large oligomers are

present in form of interlocked species, the relative simplicity of the UPLC chromatograms suggest that

they are monocyclic (as for one given macrocycle size, numerous interlocked species with different

polarity can be formed, which would substantially complicate the chromatograms). In general, the

occurrence of such large macrocycles under relatively dilute conditions (6.0 mM in 1) is unprecedented

as the production of a large number of small macrocycles is usually preferred over producing a small

number of larger entities for entropic reasons.

[41]

(although enthalpic effects due to differences in

interfacial energy also cannot be excluded). Further discussion on the topology of the macrocycles can

be found in the SI (Figure S2. 22).

In order to rationalize this unusual behavior, we performed a series of experiments to gain more

insight into the self-assembly properties of the oligomers formed from 1. First, the effect of co-solvents

was investigated. Thus, we prepared DCLs from 1, with identical building block concentration (6.0 mM)

but with increasing amounts of DMF as a cosolvent (from 10 to 90% V/V) and investigated the

(6)

47

composition of the DCLs with UPLC. The libraries were prepared by oxidizing the monomer with sodium

perborate to 85% in 30 minutes in a solvent mixture of DMF and aqueous borate buffer. The composition

of the DCLs was monitored for up to 2 months, but remained essentially unchanged after 40 days. For

detailed experimental information, see the SI Section 2.4.5.

Trimers and tetramers were always the main components in the DCLs. However, the overall LMC

content as well as the maximal detected macrocycle size decreased upon increasing DMF content

(Figure 2. 2A), giving rise finally to DCLs consisting exclusively of trimers and tetramers at high DMF

concentrations. Thus, the formation of LMCs appears to be inhibited by the presence of the organic

co-solvent, suggesting a role for hydrophobic interactions in LMC formation. We observed that in libraries

with a DMF content less than 10 % V/V and building block concentrations higher than 1 mM, occasionally,

phase separation occurred. The composition of the separated phase was however, similar to that of the

solution. For further information see the SI (Figure S2. 23). Together, these results suggested that the

LMCs are formed upon aggregation of trimers and tetramers under the given conditions.

In order to demonstrate that the formation of LMCs is a consequence of the aggregation of trimers

and tetramers, we investigated the system using a Nile Red fluorescence assay. Nile Red is a

solvatochromic dye, featuring low fluorescence intensity in aqueous solution due to aggregation, but

when incorporated into hydrophobic microenvironments it shows a significant fluorescence increase and

a characteristic blue shift of the emission maximum, as a result of encapsulation of the dye molecules

by the hydrophobic microenvironment and consequent disaggregation.

[56]

In a non-stirred, oxidized DCL

of 1 (featuring trimers, tetramers and LMCs) in aqueous buffer, Nile Red showed significantly higher

fluorescence intensity compared to that in buffer, whereas in the absence of Nile Red, neither the DCL,

nor the buffer showed fluorescence (see Figure 2. 2B).

This indicated that in aqueous buffer, the DCL contained aggregates providing a hydrophobic

microenvironment to the dye. In order to estimate the critical aggregation concentration (CAC), fully

oxidized DCLs with an increasing (0.01 - 1 mM) building block concentration (containing 230 nM Nile

Red) were prepared in aqueous buffer and the shift of the fluorescence emission maximum was

monitored. As shown in Figure 2. 2C, a sharp decrease in the fluorescence emission maximum was

detected between 0.05-0.15 mM overall building block concentration, indicating that aggregation starts

taking place in this concentration range. UPLC analyses of the samples showed that at low

concentrations, only trimers and tetramers were present, but the LMC content showed a sharp increase

in approximately the same concentration range where the Nile Red fluorescence intensity decreased

(0.1 – 0.2 mM), indicating that the formation of LMCs and aggregation are correlated.

We attempted to gain insight into the nanoscale structure of the aggregates. In fresh samples, light

microscopy showed the formation of spherical droplets with diameters between 5 and 30 μm, which form

(7)

48

larger, needle-like aggregates upon aging (see SI Figure S2. 25). Similarly, confocal fluorescence

microscopy (Nile Red staining) of a freshly prepared DCL containing LMCs showed the presence of

spherical aggregates with diameters between 5-10 μm, confirming that the trimer/tetramer aggregates

and the LMCs self-assemble into microscale objects (Figure 2. 2D). We also analyzed the samples by

transmission electron microscopy (TEM) searching for smaller nano-size assemblies, but failed to detect

any.

Figure 2. 2. A) LMC percentage (left axis) and maximal detected LMC size (n, right axis) of DCLs prepared from 1 (6.0 mM) in

50 mM borate buffer (pH = 8.2) with different amounts of DMF as a co-solvent. B) Fluorescent emission spectrum of a solution containing a DCL prepared from 1 (6.0 mM) in the absence of agitation and 230 nM Nile Red in 50 mM borate buffer (pH = 8.2, without co-solvent) at λexc = 553 nm (blue line) and the fluorescent emission spectra of Nile Red (red dots), a DCL prepared

from 1 at a total concentration of 6.0 mM (green dots) and buffer (black line) recorded under the same condition C) Fluorescence emission maximum (left axis, λexc = 553 nm ) and LMC content (right axis) of solutions containing 230 nM Nile Red and a DCL

prepared from 1 (50 mM borate buffer, pH = 8.2, without co-solvent), at different building block concentrations. Lines are drawn to guide the eye. D) Fluorescence microscopy image of an oxidized sample of building block 1 (6.0 mM in 50 mM borate buffer, pH = 8.2) containing trimers, tetramers and LMCs, stained with Nile Red (100 nM).

We interpret these results as follows: Upon oxidation and exchange, the cyclic trimer and tetramer

forms first, as observed from previous investigations of DCLs prepared from building blocks bearing the

(8)

49

same dithiol core.

[34,41,44,57]

These two amphiphilic macrocycles, featuring a hydrophobic macrocycle core

and hydrophilic tri(ethylene oxide) chains, are capable of self-assembling into supramolecular

aggregates upon exceeding a critical aggregation concentration,

[58,59]

as shown by the Nile Red

fluorescence assay. The high local concentration of disulfides in these aggregates exceeds the effective

molarity for ring closing of the smaller macrocycles, allowing LMCs to be formed. As this process takes

place mainly in a separated microphase, the energy change related to the breaking and formation of

non-covalent bonds (in this case, hydrophobic interactions which mainly arise from the interaction of the

oligomers with water molecules) is negligible. Thus, in general, the formation of these LMCs at the

covalent level is enabled by the hydrophobicity-driven self-assembly of smaller oligomers at the

non-covalent level. Upon addition of organic co-solvents, such as DMF, the solvent environment becomes

less polar and aggregation of trimers and tetramers occur to a lesser extent, resulting in a decreasing

LMC content and size in DCLs with higher co-solvent content.

Figure 2. 3. UPLC chromatogram of the isolated hexamer dissolved in MeCN:H2O 2:1 (0.1 V/V % TFA) after (A) 0 days, (B) 7 days and (C) 12 days. (D) Temporal evolution of a DCL prepared by dissolving the hexamer of 1 in MeOH (0.44 mM).

In sharp contrast to the molecular diversity observed in the non-agitated DCL prepared from 1, in a

stirred library the cyclic hexamer (1

6

) emerges exclusively, as shown in Figure 2. 1B. The hexamer

assembles as two-dimensional aggregates (vide infra) that separate from the solution as a solid

precipitate, which enabled its easy isolation by simple centrifugation and freeze-drying in 52% yield (for

a detailed procedure, see SI Section 2.4.11). We suspected that the phase separation of the hexamer is

driven by hydrophobic interactions. In order to prove this hypothesis, we assessed whether the hexamers

would disassemble again upon exposing them to organic (co)solvents. An isolated sample of the

hexamer was dissolved in a mixture of water and acetonitrile (MeCN:H

2

O 2:1 with 0.1 % TFA ) at a

(9)

50

concentration of 0.13 mM (Figure 2. 3A) and the composition of the sample was monitored with UPLC.

After 7 days, only 7 % of the DCL was present in the form of hexamers and the rest was converted to

trimers, tetramers and pentamers (Figure 2. 3B), whereas in the DCL formed after 12 days, only 4% of

hexamers were present and 29% of the library consisted of LMCs (Figure 2. 3C). Similar results were

obtained when a sample of isolated hexamer was dissolved in pure MeOH (c = 0.44 mM): in this case,

predominantly trimers and tetramers were detected at equilibrium (Figure 2. 3D). The amount of trimers

and tetramers increased parallelly (to 60 and 20 mol%, respectively), alongside with the decrease in the

amount of hexamers to 4 mol%. In this case, however, no LMCs were detectable at any stage of the

process, which is in line with our previous observations concerning the role of hydrophobic interactions

in the formation of larger oligomers. It is worth noting that under conditions which favor the formation of

LMCs (i.e. in at least partially aqueous environment), the tetramer favored over the trimer (see Figure

2. 1A and Figure 2. 3C), whereas under conditions where LMCs are not present, the trimer is the favored

species (Figure 2. 3D), which might explain that upon the dissolution of the hexamer, the trimer emerges

first (Figure 2. 3B).

Figure 2. 4. (A) Change of the product distribution with time in a DCL prepared of building block 1, showing the characteristic

sigmoidal growth of the hexamer. (B) Change of the relative concentration of the hexamer of 1 in a DCL prepared from 1 (6.0 mM) in a 9:1 mixture of aqueous borate buffer (50 mM, pH = 8.2) and DMF without seeding (squares) and upon seeding with 5.0 % (circles) and with 10 % (triangles) preformed hexamer seed at t = 0 min.

As the hexamer emerged as the sole product from a mechanically agitated DCL, we suspected it

was capable of self-replication (i.e. catalyzing its own formation driven by nanoscale self-assembly).

Thus, we monitored the change in the concentrations of the library members in time. Building block 1

was dissolved to a concentration of 2.0 mM in aqueous borate buffer and the library was stirred at 1200

rpm. The relative amount of the hexamer showed sigmoidal growth (Figure 2. 4A): a lag phase

(10)

51

(corresponding to a slow nucleation event) was followed by a rapid increase in the concentration of the

hexamer, whereupon all DCL material was converted into hexamer. Similar behavior was observed at

building block concentrations close to the CAC of the LMCs, i.e. at 0.05-0.5 mM, whereas the lag phase

increased at decreasing stirring rates (see Figure S2. 32B and C, respectively).

In order to prove the autocatalytic nature of the formation of the hexamer, we performed seeding

experiments: a pre-oxidized, non-agitated DCL prepared of 1 (6.0 mM) in a 9:1 mixture of borate buffer

(50 mM) a\nd DMF, was seeded with 5 and 10 mol% (with respect to the overall building block

concentration) pre-formed hexamer, respectively. Even in a non-stirred sample, an immediate and sharp

increase of the hexamer concentration was observed compared to the non-seeded control (Figure 2.

4B), indicating that the formation of the hexamer is autocatalytic. Additionally, we observed that in the

presence of increasing amounts of organic co-solvents both the replication rate and the final hexamer

content decreased, in line with our observations on co-solvent effects (see Figure 2. 2A and Figure S2.

32A).

Figure 2. 5. Cryogenic ((A)-(D)) and negative stain ((E)-(F)) TEM images of a stirred DCL made from pre-oxidized (80%, NaBO3) building block 1 (6.0 mM) in various stages of the self-replication process at (A) 3%; t = 0 hr (B) 4%; t = 2 hr, (C) 16%; t = 3.5 hr, (D) 18%; t = 20 hr, (E) 65%; t = 43 hr, (F) 100% hexamer content; t = 72 hr.

Based on previous examples,

[34,36,39,41,51,57]

we expected the self-replicating species to self-assemble

(11)

52

isolated hexamer did not deliver sufficiently informative data (see Figure S2. 36), we proceeded to study

the process of self-replication with cryo-TEM. A DCL at 6.0 mM building block concentration was

prepared and pre-oxidized to 80 % with sodium perborate. Stirring was continued and samples were

taken at various time points in order to monitor the dynamics of aggregate formation parallelly with UPLC

and cryo-TEM. The results are shown in Figure 2. 5. At the beginning of the monitoring process (at 3

mol% hexamer content), no nanoscale assemblies were observed (Figure 2. 5A), in line with the

previous observation that LMCs form microscale aggregates which are too large to be observed with

TEM. However, after 2 hours (4 mol% hexamer content), long, sharp-edged nanoribbons (length:

400-600 nm, width: 15-30 nm) were observed, which laterally associated into bundles (Figure 2. 5B). As the

self-replication continued (16 mol% hexamer content) the ribbons became more elongated (600-800 nm)

and more abundant (Figure 2. 5C). At 18 mol% hexamer content the single nanoribbons were not

observed anymore, and the bundles (30-40 nm wide) became the prevalent nanoscale objects (Figure

2. 5D). At a later stage of replication (65 mol% hexamer content) the bundles grew several micrometers

long and up to 80 nm wide (Figure 2. 5E). In aged samples (72 hours after the onset of the replication)

with 100 mol% hexamer content, the elongated bundles gave way to irregular platelets, with a size of ca.

100-150 nm in both directions (Figure 2. 5F).

Figure 2. 6. AFM images of a stirred DCL made from pre-oxidized (80%, NaBO3) building block 1 (6.0 mM) at (A) 54% (B) 100% conversion to hexamer (5 months old sample).

(12)

53

AFM measurements showed similar results, confirming the presence of nanoribbons and -platelets

with a constant height of 2-3 nm during the entire self-replication process (Figure 2. 6 and Figure S2.

34). Confocal fluorescence microscopy (Nile Red staining) also indicated the presence of nanoribbons

with lengths of 5-10 μm (see Figure S2. 33).

Figure 2. 7. UPLC traces of DCLs prepared from (A) 2, quickly oxidized with NaBO3 and left unagitated (B) 2, stirred for 3 days (C) 3, unagitated after 7 days (D) 3, stirred for 7 days.

These results show that the self-replication of the hexamers is concomitant with formation of

2-dimensional nanoscale assemblies. Based on the results of AFM and TEM measurements, we

hypothesize the following steps of self-assembly: Initially, the hexamer molecules pack into fibers,

featuring the hydrophobic aromatic rings at their core, which is surrounded by the hydrophilic

oligo(ethylene oxide) units (Scheme 2. 1). This hydrophobicity-driven arrangement might be further

stabilized by clustering of the side chains, which is pronounced for methyl-terminated oligo(ethylene

oxide) chains.

[60,61]

Exponential self-replication in this system is possibly a consequence of a fiber

breakage-elongation mechanism, established previously for peptide-based replicators.

[36]

Yet, unlike in

the previous systems, where the fibers were observed to elongate only longitudinally, the fibers

assembled from the hexamer are capable of stabilizing themselves by lateral association as well (Figure

2. 5B), forming nanoribbons. The reason for the different self-assembly behavior might arise from the

(13)

54

different structure of the side chains: whereas in the case of the previously reported peptide replicators,

β-sheet interactions between the peptide side chains contribute significantly to the stabilization of the

fibers, the interaction strength between the oligo(ethylene oxide) side chains is considerably smaller

compared to the energy gain resulting from the hydrophobically-driven association of the aromatic cores.

The observation that the edges of the nanoribbons are remarkably straight in the TEM and AFM images

suggests that fibers act as precursors in the formation of nanoribbons.

Figure 2. 8. Characterization of the structure (A,B) and formation (C,D) of the assemblies formed from oligomers of building

block 2 (4.0 mM in 50 mM borate buffer, pH = 8.2). A) Confocal fluorescence microscopy image of a freshly oxidized sample of building block 2 containing trimers, tetramers and LMCs, stained with Nile Red (100 nM). B) Negative stain TEM image of a stirred DCL prepared from building block 2, containing exclusively the cyclic tetramer 24. C) Change of the relative concentration

of 24 in an agitated (880 rpm) library prepared from 2. D) Change of the relative concentration of 24 in a non-agitated DCL

prepared from 2 (with 10 V/V % DMF) by fast oxidation with sodium borate (90% oxidation level) without seeding (squares) and upon seeding with 10 % (circles) preformed tetramer seed at t = 0 min.

The fact that elongated structures are produced during early stages of the assembly process

suggests that assembly at the extremities of the ribbons is initially faster than growth from the flanks of

(14)

55

the structures and also faster than their lateral association. However, as mechanical agitation is

continued and a considerable amount of the smaller macrocycles are consumed lateral association

becomes the main assembly pathway, giving rise to wider and shorter platelets. This transformation of

nanoribbons to platelets presumably represent a transition from a kinetically preferred assembly to a

thermodynamically preferred one. Note that the height of the assemblies is constant throughout the entire

replication process, i.e. the association of the nanoribbons proceeds only in one dimension.

Figure 2. 9. A) LMC percentage (left axis) and maximal detected LMC size n (right axis) of DCLs prepared from 3 (6.0 mM) in

50 mM borate buffer (pH = 8.2) with different amounts of DMF as a co-solvent. Lines are drawn to guide the eye. B) Fluorescent emission spectrum of a solution containing a DCL (total concentration 1.0 mM) prepared from 3 (in the absence of agitation) and 230 nM Nile Red in 50 mM borate buffer (pH = 8.2) at λexc = 553 nm (blue line). The fluorescent emission spectrum of Nile

Red (green), a DCL prepared from 3 at a total concentration of 1.0 mM (blue) and buffer (black) recorded under the same conditions is also indicated. C) Fluorescent emission maximum (left axis, λexc = 553 nm ) and LMC content (right axis) of

solutions containing 240 nM Nile Red and a pre-oxidized DCL prepared from 3 (50 mM borate buffer, pH = 8.2), at different building block concentrations. D) Particle size distribution in a non-agitated DCL prepared from building block 3 (6.0 mM) in aqueous borate buffer (50 mM, pH = 8.2), measured with dynamic light scattering.

(15)

56

We also considered the possibility that Na

+

ions might act as templates by forming crown-ether like

chelates upon interactions with the oligo(ethylene oxide) side chains of neighboring molecules of the

hexamer. However, the replication rate shows no readily interpretable dependence on the concentration

of Na

+

in the 0-50 mM range (see Figure S2. 32D). We were also interested to what extent the molecular

structure of the monomer affects the dual assembly modes described above. Thus, we synthesized two

analogues of 1 containing one ethylene oxide unit less (2) and more (3) than 1. Analogue 2 oxidized only

very slowly when exposed to air in the absence of stirring, possibly due to its low solubility in aqueous

borate buffer. However, quick oxidation with perborate in the absence of agitation gave rise to a DCL

containing mainly trimers and tetramers but also a significant amount of LMCs (Figure 2. 7A).

Fluorescence microscopy showed spherical aggregates (Figure 2. 8A) which were similar to those

observed for 1, whereas with TEM no nanoscale structures were detectable (Figure S2. 27). Addition of

large amounts of co-solvent led to the disappearance of LMCs, analogously to 1 (Figure S2. 26). When

stirred, the cyclic tetramer emerged as the only product (Figure 2. 7B). Fluorescence microscopy and

TEM showed the presence of nanoribbons (Figure S2. 33). Similarly to the hexamer formed from 1, this

tetramer could also be easily isolated in 76 % yield. Compared to the hexamer of 1, the tetramer of 2

showed a weak autocatalytic effect; more precise analysis of the seeding process was hampered by

analytical difficulties related to the poor solubility of the tetramer of 2 (see Figure 2. 8C, D). Assembly of

2 occurs for a smaller macrocycle size than for 1 which can be rationalized based on the fact that building

block 2 is more hydrophobic than 1. Therefore fewer units of 1 are required to generate a sufficient

hydrophobic driving force to enable self-assembly.

[41]

In sharp contrast, 3 does not show preference for any specific macrocycle and, regardless of

mechanical agitation, gives rise to LMCs up to 55

mer

upon oxidation (Figure 2. 7C-D and Figure S2. 29).

The mechanism of the formation of the LMCs for 3 resembles that for 1, as shown by co-solvent (Figure

2. 9A) and fluorescence (Figure 2. 9B-C) experiments. TEM shows no detectable aggregates (see

Figure S2. 30A), whereas in fluorescence microscopy, spherical aggregates similar to those observed in

case of 1, were detectable (see Figure S2. 30B). Dynamic light scattering indicated the presence of

aggregates with a diameter of ca. 200 nm (Figure 2. 9D). The presented data for DCLs formed from 1

support a complex self-assembly energy landscape (Scheme 2. 2), where mechanically triggered

autocatalysis allows to access specific assembly modes by lowering activation barriers. In the absence

of mechanical agitation, the oxidized monomers first form trimers and tetramers, which self-assemble

into less defined spherical aggregates. The corresponding part of the (simplified) energy landscape can

be represented by a wide and relatively shallow energy minimum, with several local minima,

corresponding to the trimer/tetramer aggregates and LMCs, whose mutual interconversion reactions

feature low activation barriers. However, among the aggregates formed, also the hexamer assemblies,

(16)

57

capable of autocatalytic growth, are present. Nevertheless, these small aggregates (primary nuclei)

require a very long time to grow as the number of autocatalytic fiber ends is negligible. This implies a

high energy barrier toward the formation of the thermodynamically more stable hexamer nanoribbons.

Upon stirring, however, mechanical energy is administered to the system, resulting in the breakage of

the primary nuclei. As a result of this process, the number of free fiber ends growth rapidly (potentially

exponentially

36

), which allows for the autocatalytic growth to set in. The primary hexamer assemblies

(fibers) serve as a template for the formation of further hexamer molecules, either at their ends (resulting

in longer fibers) or at their side (leading to nanoribbons). In other words, mechanical energy supply

lowers the activation barrier of hexamer formation by enabling an autocatalytic pathway. DCLs made

from building block 2 follow very similar assembly paths compared to DCLs made from 1, except that

now the tetramer and not the hexamer is the species that assembles into fibers and ribbons.

Scheme 2. 2. Simplified Potential Energy Landscape of Dynamic Combinatorial Libraries of Building Block 1 in water.

2.3. Conclusions

In conclusion we showed for the first time in the context of dynamic combinatorial chemistry that a

single building block can give rise to two systems featuring remarkably different modes of self-assembly.

Without agitation, self-assembly of cyclic trimers and tetramers into less-defined aggregates and

subsequent disulfide exchange leads to a diverse mixture of unprecedentedly large covalent

(17)

58

macrocycles. With agitation one specific macrocycle self-assembling into well-defined nanoribbons and

–platelets, forms in an autocatalytic manner, enabled again by disulfide exchange. Thus, due to the

presence of dynamic covalent bonds, the difference in self-assembly modes at the non-covalent level is

also reflected at the covalent level. The fact that aggregation is accompanied by disulfide exchange only

in the former case is most likely a result of the higher thermodynamic stability of the hexamer assemblies

(due to the close packing of hexamer units) compared to the ill-defined aggregates of trimers and

tetramers. Systems that may be channeled into distinct self-assembly pathways are receiving increasing

attention in nanotechnology and materials science in the last years.

[62–67]

Extension of these systems to

incorporate a dynamic covalent level, as shown now in our work, opens the way to multifaceted dynamic

self-assembling systems of potential interest in the context of materials science and artificial life.

(18)

59

2.4. Supporting Information

2.4.1. General remarks

All chemicals, unless otherwise stated, were purchased from Sigma-Aldrich and used as received. Acetonitrile (ULC-MS grade), water (ULC-MS grade) and trifluoroacetic acid (HPLC grade) were purchased from Biosolve BV. Reagent grade solvents were purchased from Macron and used as received. Anhydrous solvents used in synthesis were freshly collected from a dry solvent purification system prior to use. Flash column chromatography was performed on a Reveleris ® X2 Flash Chromatography System (Grace Davison Discovery Sciences, Deerfield IL) on normal phase silica cartridges. NMR spectra were taken using a Varian VXR400 spectrometer. HRMS spectra were recorded on a LTQ Orbitrap XL instrument in ESI ionization mode. Fluorescence measurements were performed on a JASCO FP 6200 fluorimeter using quartz cuvettes with 1 cm path length at an excitation wavelength of 553 nm.

Sample preparation

Generally, samples were diluted to 100 μM with a 7: 3 V:V mixture of water (UPLC grade) and DMF (HPLC grade), containing 0.1 V/V% TFA (HPLC grade) and 5 μL of this diluted sample was injected for UPLC measurements. In experiments studying the co-solvent dependence of the composition of the DCLs (Figure 2. 2A), 1 μL of undiluted samples was injected. In experiments studying the concentration dependence of LMC content (Figure 2. 2B), 10 μL of undiluted samples were injected. In experiments studying the disassembly of 16 in methanol, 2.5 μL of undiluted samples were injected.

UPLC analysis

UPLC measurements were performed on a Waters Acquity H-class system equipped with a PDA detector, at a detection wavelength of 254 nm. UPLC analyses, unless otherwise stated, were performed on an Aeris WIDEPORE 3.6 μm XB-C18 (150 × 2.1 mm) column, purchased from Phenomenex, using ULC-MS grade water (eluent A) and ULC-MS grade acetonitrile (eluent B), containing 0.1 V/V % TFA as a modifier. A flow rate of 0.3 mL/min and a column temperature of 35 °C were applied.

Methods for the analysis of DCLs prepared from 1 and 2: t / min % A 0 60 7 38 10 34 17 5 18 5 19 60 20 60 Method for the analysis of DCLs prepared from 3:

t / min % A 0 60 17 25 18 25 19 60 20 60 UPLC-MS analysis

UPLC-MS measurements were performed using a Waters Acquity UPLC H-class system coupled to a Waters Xevo-G2 TOF. The mass spectrometer was operated in the positive electrospray ionization mode with the following ionization parameters: capillary voltage: 3 kV, sampling cone voltage: 20 V, extraction cone voltage: 4 V, source gas temperature: 120°C, desolvation gas temperature: 450°C, cone gas flow (nitrogen): 1 L/h, desolvation gas flow (nitrogen): 800 L/h.

Negative-staining Transmission Electron Microscopy

Samples were diluted to 60-fold using UPLC grade water. A small drop (5 µL) of sample was then deposited on a 400 mesh copper grid covered with a thin carbon film (supplied by Agar Scientific). After 30 seconds, the droplet was blotted on filter paper. The sample was then stained with a solution of 2% uranyl acetate (4 µL) deposited on the grid, subsequently washed and blotted on filter paper after 30 seconds. The staining procedure was repeated a second time, this time without the washing and

(19)

60

blotting step. The grids were observed in a Philips CM12 electron microscope operating at 120 kV. Images were recorded on a slow scan CCD camera.

Cryo Transmission Electron Microscopy

A 10 μL drop of the sample was placed on a Quantifoil 3.5/1 holey carbon coated grid. Blotting and vitrification in ethane was done in a Vitrobot (FEI, Eindhoven, the Netherlands). The grids were observed in a Tecnai T20 cryo-electron microscope operating at 200 keV with a Gatan model 626 cryo-stage. Images were recorded under low-dose conditions with a slow-scan CCD camera.

Atomic Force Microscopy

AFM samples were prepared by depositing 100 µL of the sample (diluted to 10 µM building block concentration with UPLC grade water) onto a clean mica surface (Grade V1, Van Loenen Instruments). Subsequently, the solvent was evaporated in a gentle stream of air in ca. 20 minutes. The surface was then washed with 100 µL of UPLC grade water and blotted into a piece of paper twice and finally air-dried. The AFM measurements have been performed using a Bruker Multimode 8 instrument in Scan Asyst-Air imaging mode. Measurements were performed in air at room temperature. As a probe, a ScanAsyst Air (Bruker) silicon tip on a nitride cantilever was used with the following parameters: length: 115 µm, width: 25 µm, resonance frequency: 70 kHz, force constant: 0.4 N/m. The images were recorded with frequencies between 0.5 and 1.5 Hz and analyzed with NanoScope Analysis 1.50 software (Bruker Corporation, 2015).

Optical Microscopy

Optical microscopy was performed on a Nikon Eclipse TS 100 instrument. Length calibration was performed by recording the image of a 40 μm scale bar with the same settings as for the corresponding image. The images were processed using ImageJ software.

Confocal Fluorescence Microscopy

Confocal fluorescence microscopy measurements were performed on a MicroTime 200 instrument (PicoQuant, Berlin, Germany), using an excitation wavelength of 532 nm (the excitation power was set at 4 μW), with a 542 nm long-pass edge emission filter (BLPO1-532R, Semrock, New York). For the measurement, 50 µL sample was deposited on the microscope slide and 5.0 µL of Nile Red (1.0 µM solution in methanol) was added. The measurements were performed after an incubation time of 2 minutes.

Dynamic Light Scattering

Dynamic light scattering measurements were performed on a NanoBrook 90Plus PALS Particle Size Analyzer (Brookhaven, NY), using a 659 nm laser at 90º detection angle. Samples were prepared using 50 mM aqueous borate buffer (pH = 8.2), previously filtered with a 0.2 μm pore size filter. For each measurement, 10 parallels were recorded.

2.4.2. Synthetic Procedures

To a vigorously stirred, ice-cooled solution of the monomethylated oligoethylene glycol (50 mmol) and triethylamine (8.5 mL, 61 mmol, 1.2 eq) in anhydrous DCM (60 mL) a solution of methylsulfonyl chloride (5.4 mL, 70 mmol, 1.4 eq) was added dropwise. The reaction mixture was stirred overnight at 0-5°C in a cold room. The mixture was warmed to room temperature. DCM (20 mL) and deionized water (50 mL) was added and the solution stirred for 30 minutes to obtain a clear biphasic mixture. The layers were separated and the organic layer was washed with 1M HCl (540 mL) and water (500 mL). The combined aqueous layers were back-extracted with DCM (5 ×100 mL). The combined organic layers were dried over Na2SO4 and concentrated.

The obtained crude mesylate was dissolved in a mixture of 25 w/w% NH3 in water (600 mL) and NH4Cl (90 g) and stirred for 48 h. NaCl (90 g) was dissolved in the reaction mixture, which was extracted with DCM (3 ×500 mL). The combined organic layers were dried over Na2SO4 and concentrated to obtain the corresponding amine as a yellow oil, which was used without

further purification in the following reactions.

S2a. Yield: 4.05 g (67%). 1H NMR (400 MHz, CDCl3) δ 3.35 (dtt, J = 4.6, 2.2, 1.1 Hz, 2H), 3.28 (dddd, J = 5.1, 4.0, 2.2, 1.1

(20)

61

69.95, 58.70, 58.67, 41.53. ESI-HRMS: [M+H] + found: 120.1017 Da (expected: 120.1019 Da). Analytical data matched that

reported previously.[68]

S2b. Yield: 7.24 g (88%). 1H NMR (400 MHz, CDCl3) δ 5.26 (s, 2H), 3.58 (th, J = 5.6, 2.6 Hz, 6H), 3.50 – 3.46 (m, 2H),

3.44 (t, J = 5.2 Hz, 2H), 3.31 (s, 3H), 2.79 (t, J = 5.2 Hz, 2H), 1.44 (s, 2H). 13C NMR (101 MHz, CDCl3) δ 73.29, 71.79, 70.45, 70.13, 58.84, 53.41, 41.66. ESI-HRMS: [M+H] + found: 164.1280 Da (expected: 164.1281 Da). Analytical data matched that

reported previously.[68]

S2c. Yield: 8.65 g (84%). 1H NMR (400 MHz, CDCl3) δ 3.66 – 3.50 (m, 10H), 3.48 (dd, J = 5.7, 3.6 Hz, 2H), 3.44 (t, J = 5.2

Hz, 2H), 3.31 (s, 3H), 2.79 (t, J = 5.2 Hz, 2H), 1.46 (s, 2H). 13C NMR (101 MHz CDCl3) δ 73.30, 71.82, 70.50, 70.47, 70.46,

70.40, 70.17, 58.89, 41.70. ESI-HRMS: [M+H] + found: 208.1542 Da (expected: 208.1543 Da). Analytical data matched that

reported previously.[69]

To an ice-cooled suspension of 3.5-bis(trithyltio)benzoic acid S3[34] (1.2 g, 1.8 mmol) and HOBt (0.364 g, 2.4 mmol, 1.3

eq) in anhydrous DCM (60 mL) was added EDC.HCl (0.934 g, 4.9 mmol, 2.7 eq). A clear yellow solution was obtained.

Subsequently, an ice-cooled solution of the corresponding amine S2a-S2c (2.7 mmol, 1.5 eq) and diisopropylethylamine (3.1 mL, 18 mmol, 10 eq) in anhydrous DCM (30 mL) was added dropwise. The mixture was allowed to warm to room temperature and stirred overnight. DCM (50 mL) was added and the solution was washed with 1M HCl (30 mL), saturated NaHCO3 solution

(30 mL) and water (30 mL). The organic phase was dried over MgSO4 and concentrated in vacuo. The residue was dissolved

in DCM (4 mL) and purified by flash column chromatography to obtain the product as a white foamy solid.

S4a: Yield: 83% (1.14 g). Gradient for flash column chromatography (24 g silica): 0 min: 100% DCM, 0-2 min: 100% DCM

to 100% hexanes, 2-3 min: 100% hexanes, 3-13 min: 100% hexanes to 50% hexanes in EtOAc, 13-15 min: 50% hexanes in EtOAc, 15-17 min : 50% hexanes in EtOAc to 60% hexanes in EtOAc, 17-18 min: 60% hexanes in EtOAc, 18-20 min : 60% hexanes in EtOAc to 100% EtOAc, 20-27 min: 100% EtOAc. 1H NMR (400 MHz, CDCl3) δ 7.39 – 7.30 (m, 12H), 7.21 (m, 18H),

7.12 (t, J = 1.7 Hz, 1H), 6.88 (d, J = 1.6 Hz, 2H), 3.63 (dd, J = 5.8, 2.9 Hz, 2H), 3.59 – 3.52 (m, 4H), 3.47 (q, J = 5.4 Hz, 2H), 3.39 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 165.96, 144.14, 142.29, 134.80, 134.05, 132.40, 130.19, 129.92, 127.81, 127.58,

126.81, 77.45, 77.13, 76.81, 71.90, 71.34, 70.32, 69.90, 60.40, 59.09, 39.63, 14.25. ESI-HRMS: [M+H] + found: 772.2914 Da

(expected: 772.2914 Da), [M+Na] + found: 794.2740 Da (expected: 794.2733 Da).

S4b: Yield: 47% (0.715 g, 1.2 g scale), 67 % (2.303 g, 3 g scale). Gradient for flash column chromatography (24 g silica):

0 min: 100% DCM, 0-1 min: 100% DCM to 100% hexanes, 2-4 min: 100% hexanes, 4-18 min: 100% hexanes to 50% hexanes in EtOAc, 18-19 min: 50% hexanes in EtOAc, 19-21 min : 50% hexanes in EtOAc to 100% EtOAc, 21-29 min: 100% EtOAc. 1H

NMR (400 MHz, CDCl3) δ 7.23 (dt, J = 5.4, 3.3 Hz, 12H), 7.14 – 7.05 (m, 18H), 7.01 (t, J = 1.6 Hz, 1H), 6.77 (d, J = 1.6 Hz, 2H), 5.66 (t, J = 5.4 Hz, 1H), 3.60 – 3.52 (m, 6H), 3.47 – 3.41 (m, 4H), 3.35 (dd, J = 10.1, 5.0 Hz, 2H), 3.22 (s, 3H), 1.95 (s, 1H).

13C NMR (101 MHz, CDCl3) δ 165.92, 144.12, 142.27, 134.76, 134.04, 132.41, 129.89, 127.79, 126.79, 71.95, 71.31, 70.60,

70.57, 70.37, 69.89, 59.02, 39.65. ESI-HRMS: [M+H]+ found: 816.3168 Da (expected: 816.3176 Da). [M+Na]+ found: 838.2992

Da (expected: 838.2995 Da).

S4c: Yield: 55% (0.852 g). Gradient for flash column chromatography (24 g silica): 0 min: 100% DCM, 0-2 min: 100% DCM

to 100% hexanes, 2-3 min: 100% hexanes, 3-19 min: 100% hexanes to 100% EtOAc, 19-21 min: 100% EtOAc.1H NMR (400

MHz, CDCl3) δ 7.38 – 7.29 (m, 12H), 7.20 (dd, m, 18H), 7.11 (t, J = 1.7 Hz, 1H), 6.88 (d, J = 1.7 Hz, 2H), 5.82 (t, J = 5.4 Hz, 1H), 3.71 – 3.57 (m, 10H), 3.57 – 3.49 (m, 4H), 3.46 (q, J = 5.0 Hz, 2H), 3.35 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 165.92,

144.12, 142.27, 134.74, 134.03, 132.44, 129.89, 127.79, 126.79, 71.90, 71.32, 70.63, 70.60, 70.51, 70.37, 69.91, 59.02, 39.67. ESI-HRMS: [M+H]+ found: 860.3438 Da (expected: 860.3438 Da).

(21)

62

9 mL TFA was degassed through 3 freeze-pump-thaw cycles in a 25-mL Schlenk tube and added to a 50-mL round-bottomed flask containing protected amide S4a-S4c (1.4 mmol) under nitrogen. The orange solution was sonicated for 1 minute and stirred at room temperature for 15 minutes. Subsequently, Et3SiH (1.0 mL, 6.3 mmol, 4.5 eq) was added dropwise and the

solution stirred for further 15 minutes. TFA was removed in vacuo via the Schlenk line and condensed into a cold trap immersed in liquid nitrogen. The remaining suspension was dissolved in 70 mL of a degassed mixture of MeOH-H2O (9:1, 70 mL) and

washed with 3*100 mL heptane. Methanol was removed in vacuo at 40°C and the remaining suspension was freeze-dried for 36 hours to furnish the corresponding building block 1-3.

1: Sticky, yellow solid. Yield: 90% (0.416 g). 1H NMR (400 MHz, (CD

3)2SO) δ 8.47 (t, J = 5.6 Hz, 1H), 7.46 (d, J = 1.7 Hz,

2H), 7.36 (t, J = 1.7 Hz, 1H), 5.67 (s, 2H), 3.48 (ddd, J = 7.1, 3.1, 1.5 Hz, 8H), 3.40 – 3.32 (m, 4H), 3.20 (s, 3H). 13C NMR (101

MHz, CD3OD) δ 167.55, 135.77, 133.83, 130.38, 123.68, 71.48, 70.10, 69.94, 69.85, 68.98, 57.66, 39.60. ESI-HRMS: [M+H]+

found: 332.0985 Da (expected: 332.0985 Da), [M+Na]+ found: 354.0802 Da (expected: 354.0810 Da).

2: White-yellow solid. Yield: 98% (0.393 g). 1H NMR (400 MHz, DMSO-d6) δ 8.49 (t, J = 5.6 Hz, 1H), 7.46 (d, J = 1.7 Hz,

2H), 7.36 (t, J = 1.7 Hz, 1H), 5.70 (s, 2H), 3.52 – 3.45 (m, 4H), 3.44 – 3.39 (m, 2H), 3.36 (t, J = 5.8 Hz, 2H), 3.21 (s, 3H). 13C

NMR (101 MHz, CD3OD) δ 167.60, 135.75, 133.84, 130.35, 123.64, 71.52, 69.60, 69.03, 57.77, 39.55. ESI-HRMS: [M+Na]+

found: 310.0543 Da (expected: 310.0542 Da).

3: Yellow oil. Yield: 93% (0.488 g). 1H NMR (400 MHz, (CD3)2SO) δ 8.47 (s, 1H), 7.46 (d, J = 1.5 Hz, 2H), 7.36 (s, 1H),

5.67 (s, 2H), 3.48 (dd, J = 13.3, 4.8 Hz, 12H), 3.42 – 3.31 (m, 4H), 3.20 (s, 3H). 13C NMR (101 MHz, CD3OD) δ 167.47, 135.75,

133.83, 130.40, 123.75, 71.51, 70.12, 70.07, 69.91, 69.88, 69.02, 57.70, 39.67. ESI-HRMS: [M+H]+ found: 376.1246 Da (expected: 376.1247 Da), [M+Na]+ found: 398.1062 Da (expected: 398.1066 Da).

(22)

63

2.4.3. NMR Spectra

Figure S2. 1. 1H NMR (400 MHz, CDCl3) spectrum of S2a.

(23)

64

Figure S2. 3. 1H NMR (400 MHz, CDCl3) spectrum of S2b.

(24)

65

Figure S2. 5. 1H NMR (400 MHz, CDCl3) spectrum of S2c.

(25)

66

Figure S2. 7. 1H NMR (400 MHz, CDCl3) spectrum of S4a.

(26)

67

Figure S2. 9. 1H NMR (400 MHz, CDCl3) spectrum of S4b.

(27)

68

Figure S2. 11. 1H NMR (400 MHz, CDCl3) spectrum of S4c.

(28)

69

Figure S2. 13. 1H NMR (400 MHz, d

6-DMSO) spectrum of 1.

(29)

70

Figure S2. 15. 1H NMR (400 MHz, d6-DMSO) spectrum of 2.

(30)

71

Figure S2. 17. 1H NMR (400 MHz, d6-DMSO) spectrum of 3.

Figure S2. 18. 13C NMR (101 MHz, d

(31)

72

2.4.4. UPLC chromatograms of monomers 1, 2, and 3.

Figure S2. 19. UPLC chromatogram of 1 (BEH C18 column, 5 to 95 % MeCN in H2O, detection wavelength 254 nm.)

Figure S2. 20. UPLC chromatogram of 2 (BEH C18 column, 5 to 95 % MeCN in H2O, detection wavelength 254 nm.)

Figure S2. 21. UPLC chromatogram of 3 (BEH C18 column, 5 to 95 % MeCN in H2O, detection wavelength 254 nm.

2.4.5. Procedure for preparation of DCLs with perborate-mediated oxidation

The following shows a typical example. A DCL made from 1 (6.0 mM), pre-oxidized with NaBO3 to 80% in borate buffer in

the presence of 20 V/V % DMF, was prepared in the following manner: 1. 4.03 mg NaBO3.4H2O (26.2 μmol) was weighed in a vial.

2. As the overall building block concentration is 6 mM, a 60 mM solution was prepared in borate buffer. This required

26.2 μmol

60 μmol/mL = 437 μL borate buffer.

3. 437 μL borate buffer was added to the vial containing NaBO3.4H2O and the solution was sonicated for 2 minutes until

the entire amount of salt was dissolved.

4. 2.62 mg 1 (7.92 μmol) is weighed in a 2 mL vial containing a small stirring bar. 5. This required a total solvent volume of 7.92 μmol

6 μmol/mL = 1320 μL solvent.

6. This required 1320 ×20% = 264 μL DMF. 7. This required a total amount of 1320 μL ×6 mM × 80%

60 𝑚𝑀 = 106 μL NaBO3 solution.

8. To fill up the remaining volume, 1320 – 264 -106 = 950 μL buffer was required.

9. Thus, 264 μL DMF was added to the vial containing the building block (with a Hamilton syringe). 10. Subsequently, 950 μL buffer was added (with a Eppendorf pipette).

(32)

73

11. The oxidant was added in portions (to avoid over-oxidation of the thiol groups). 2.5 - 3% oxidant was added at once safely in 1-minute intervals without the risk of over-oxidation. I.e. 80 % of oxidant was added in 80 %

2.5% / 𝑚𝑖𝑛= 32 portions

of 106 μL

32 = 3.31 μL, adding one portion in every minute, under vigorous stirring (1200 rpm).

12. After adding the last portion of oxidant, stirring was discontinued. According to our experience, no replication happened during this period and the composition of the libraries prepared in this manner was similar to those prepared by oxidation with air in the absence of stirring.

13. The composition of the resulting DCL was monitored with UPLC over the course of 7-40 days.

2.4.6. LC-MS analyses of DCLs

From environmental reasons, the large number of LC-MS spectra are not published in the printed version of this thesis. However, they are accessible free of charge in the online Supporting Information of the original publication.

2.4.7. General remark on the topology of LMCs based on the mass spectra

For the large oligomers, UPLC-TOF-MS can in principle distinguish between cyclic and linear species as the latter ones would contain two extra hydrogen atoms at the chain ends. In practice, as the amount of the corresponding oligomers (and thus the corresponding signal intensities and signal-to-noise ratios) decrease with increasing oligomer length, we could only confirm the presence of cyclic oligomers up to 35mers, based only on the MS analysis. Furthermore, the low abundance hampered the MS detection of oligomers above 38-mer (building block 1) and 42-mer (building block 3), respectively. Additional peaks most likely corresponding to even larger oligomers were clearly observed in the UPLC chromatograms (see Figure S2. 24and Figure S2. 29).

If linear oligomers were present in considerable amount, thiols and linear dimers would also be detectable in similar concentrations, due to thiol-disulfide exchange. UPLC analysis of fully oxidized samples show that the amount of these free thiol-containing species are below the detection limit. One can argue that for larger oligomers, the linear and cyclic species composed of the same number of monomer units cannot be separated by UPLC because the difference in polarity conferred by the presence of two thiol groups is relatively small due to the large size of both species. Thus, we performed the following experiment: For each building block 1-3, a UPLC spectrum of a fully oxidized library (6 mM), containing considerable amounts of LMCs, alongside with a low amount of organic co-solvent, was recorded. Subsequently, a 100 μL portion of each sample was diluted with 900 μL isopropanol, resulting in DCLs (1.2 mM) containing 90 V/V % organic co-solvent. As discussed in detail in the main text, these libraries contained only neglectable amounts of LMCs as the hydrophobicity-driven aggregation of trimers and tetramers, which is necessary for the formation of large oligomers, is far less pronounced in mainly organic media (see

Figure 2. 2in the Main Text).

If there were a considerable amount of linear large oligomers, containing free thiol groups, these would cause a considerable increase in the amount of free thiols (monomer and linear dimer). However, the linear thiol concentration increased only by a negligible amount (0.1 % and 1.1 % for building block 1 and 3, respectively, whereas in the case of building block 2, no change was detectable), as shown in Figure S2. 22. Thus, based on our analytical data, the detected large oligomers are most probably macrocyclic species, with at most a few percent of linear species present.

Figure S2. 22. Relative concentration of free thiol (calculated as c(SH) = c(monomer)+0.5×c(dimer); black bars) and LMC (red bars) content of fully-oxidized, non-stirred DCLs (6 mM) prepared from building block A) 2, B) 1, C) 3 before (left-side bars) and after (right-side bars) diluting into 90 V/V% isopropanol.

(33)

74

2.4.8. Additional data on the LMCs from building block 1

Remark: In oxidized, non-stirred DCLs prepared from 1, stable colloids are formed at concentrations below 1 mM. In

contrast, in libraries with a DMF content less than 10 V/V % and building block concentration higher than 1 mM, phase separation occurred This process was also evident from UPLC analysis of the samples, showing extensive peak area loss for samples that exhibited phase separation (Figure S2. 23C). After re-dissolving phase-separated material in MeCN, UPLC analysis revealed that it had a composition similar to that of the solution phase, consisting mainly of trimers and tetramers (Figure S2. 23A). Consistently, upon the addition of increasing amounts of co-solvent, the phase-separated material could be re-dissolved as evidenced by the increase in the overall UPLC peak area (Figure S2. 23B).

Figure S2. 23. The effect of organic co-solvents on the solubility of LMCs prepared from 1. A sample prepared by isolating the oily substrate (see inset, arrow pointing towards an oil droplet), formed in a vial containing a 2 months old non-agitated DCL prepared from 1. The oil was dissolved in MeCN and UPLC chromatograms were recorded by adding A) 0% B) 10 % DMSO to the sample. Note that the intensity scales of the two chromatograms are the same. C) Total UPLC peak area of a DCL prepared from 1 (6.0 mM) with different amounts of DMF as a co-solvent.

Figure S2. 24. Partial UPLC chromatogram of a DCL prepared from 1, showing peaks most likely corresponding to oligomers

(34)

75

Figure S2. 25. Optical microscopy images taken of a non-agitated DCL prepared of 1 (6.0 mM) in borate buffer (50 mM, pH = 8.2), pre-oxidized 80 % with sodium perborate (A) 1 hour (B) 1 day (C) 4 months after preparation. Note the presence of spherical aggregates in the fresh samples.

2.4.9. Additional characterization of the LMCs formed from building block 2.

Figure S2. 26. UPLC chromatograms of a DCL prepared from building block 2 (4 mM), oxidized to 95% with sodium perborate in aqueous borate buffer (50 mM, pH = 8.2) containing A) 5 V/V % DMSO B) 75 V/V % MeCN.

(35)

76

Figure S2. 27. Negative staining TEM image of a DCL prepared from building block 2 (4 mM) in borate buffer (50 mM, pH = 8.2), oxidized to 90% with sodium perborate. According to the UPLC analysis, the sample contained trimers, tetramers and large macrocycles, whereas TEM shows no indication for the presence of nanoscale objects.

2.4.10. Additional data on the LMCs from building block 3.

Figure S2. 28. A sample prepared by injecting a 2 months old, non-agitated DCL prepared from 3 in aqueous borate buffer (50 mM, pH = 8.2) after adding A) 0% B) 10 % DMSO to the sample. Note that the intensity scales of the two chromatograms are the same.

(36)

77

Figure S2. 29. Partial UPLC chromatogram of a DCL prepared from 3, showing peaks most likely corresponding to oligomers

340-351 (not detected in LC-MS due to low signal intensity).

Figure S2. 30. Microscopic analyses of a DCL prepared from building block 3 (6 mM) in borate buffer (50 mM, pH = 8.2), fully oxidized by air. A) Negative staining TEM image image B) Confocal fluorescence microscopy image, stained with Nile Red (100 nM).

2.4.11. Additional characterization of hexamer 1

6

(37)

78

1 (7.7 mg, 23 μmol) was suspended in aqueous sodium borate buffer (1.2 mL, 50 mM, pH = 8.2) and was left standing for

1 day. Subsequently, it was stirred for one day. At this time, UPLC analysis showed complete conversion of the monomer to hexamer. The resulting suspension was centrifuged (17 min, 3900 rpm). The pellet was washed with doubly distilled water (2×1200 μL) to remove excess sodium borate and centrifuged. The residue was freeze-dried overnight to yield 16 (4.0 mg, 2.0

μmol) as a white solid in 52 % yield. The purity of the product was confirmed by UPLC (Figure S2. 31).

Figure S2. 31. UPLC chromatogram of isolated 16, freshly dissolved in MeCN.

Figure S2. 32. The effect of various physical-chemical parameters on the replication kinetics of 16. Time evolution of DCLs

(38)

79

pre-oxidized with NaBO3) B) at different concentrations ([1] = 0.05, 0.1 and 0.5 mM) C) at different stirring rates (100, 600 and

1200 rpm) at [1] = 2.0 mM. D) Change of the relative concentration of 16 the in a DCL prepared from 1 (6.0 mM) in UPLC grade

water with different concentrations of NaCl. Note that replication is also relatively fast in the absence of Na+ ions.

Figure S2. 33. Confocal fluorescence microscopy image of a sample containing exclusively 16 (6.0 mM) in borate buffer (50

mM, pH = 8.2) stained with Nile Red (100 nM).

Figure S2. 34. AFM images of a library prepared from 1 (6.0 mM) by oxidizing it to 80 % with NaBO3 in aqueous borate buffer

and subsequently stirred for A) 8.5 hours (early stage of self-assembly) and B) 12 hours (late stage of self-assembly). Note that at the early stage of self-assembly aggregates of 4-5 nanoribbons (150 – 300 nm long) are observed, whereas at the late stage of self-assembly several tens of nanoribbons self-assemble to give large aggregates (more than 6 µm long).

(39)

80

Figure S2. 35. TEM images taken of samples from of a DCL prepared from 1 (6.0 mM) in borate buffer (50 mM, pH = 8.2), pre-oxidized 80 % with sodium perborate, after stirring for A) 0 hours , B) 2 hours, C) 3.5 hours, D) 17 hours, E) 20 hours, F) 72 hours. Note the absence of nanoribbons at 0-1 hour, the presence of increasingly elongated and laterally associated nanoribbons (associated with the self-replication and self-assembly of 16) from 2-43 hours, as well as the presence of

nanoplatelets at 72 hours.

(40)

81

2.4.12. Additional characterization of tetramer 2

4 Synthesis of 24 from 2:

2 (5.2 mg, 15 μmol) was suspended in aqueous sodium borate buffer (2.74 mL, 50 mM, pH = 8.2). Sodium perborate

(60mM, 304 μL in sodium borate buffer) was added in portions of 12.2 μL over the course of 25 minutes. Subsequently, the reaction mixture was stirred for 5 days until UPLC analysis showed complete conversion of the monomer to the tetramer. The resulting suspension was centrifuged (5 min, 5000 rpm). The pellet was washed with doubly distilled water (1 mL) to remove excess sodium borate and centrifuged. The residue was freeze-dried overnight to yield 24 (4.0 mg, 3.5 μmol) as a white solid in

76 % yield. The purity of the product was confirmed by UPLC (Figure S2. 37).

Figure S2. 37. UPLC chromatogram of isolated 24, freshly dissolved in 2-propanol.

Figure S2. 38. Confocal fluorescence microscopy image of a sample containing exclusively 24 (3.84 mM) in borate buffer (50

(41)

82

2.5. Author Contributions

D.K. and S. O. conceived the project. D. K. , S. O. and M. C. A. S. designed the experiments. D. K.

performed synthesis and characterization of new compounds, DCl preparation, UPLC, LC-MS and

spectroscopic analyses. M. C. A. S. performed cryo-TEM measurements, M. T. performed negative

staining TEM measurements, G. M. S. performed AFM measurements, V. V. K. performed fluorescence

microscopy imaging. D. K. analyzed data and D. K. and S. O. wrote the manuscript. Jacob Baas is

gratefully acknowledged for the PXRD measurements. Gaël Schaeffer is gratefully acknowledged for

fruitful discussions.

2.6. References

[1] P. Nissen, J. Hansen, N. Ban, P. B. Moore, T. A. Steitz, Science 2000, 289, 920–930. [2] T. Suganuma, J. L. Workman, Annu. Rev. Biochem. 2011, 80, 473–499.

[3] K. Meguellati, S. Ladame, Top. Curr. Chem. 2012, 322, 291–314.

[4] Y. Jin, C. Yu, R. J. Denman, W. Zhang, Chem. Soc. Rev. 2013, 42, 6634–6654. [5] Y. Jin, Q. Wang, P. Taynton, W. Zhang, Acc. Chem. Res. 2014, 47, 1575–1586.

[6] P. T. Corbett, J. Leclaire, L. Vial, K. R. West, J.-L. Wietor, J. K. M. Sanders, S. Otto, Chem. Rev. 2006, 106, 3652–3711. [7] S. Otto, Acc. Chem. Res. 2012, 45, 2200–2210.

[8] F. B. L. Cougnon, J. K. M. Sanders, Acc. Chem. Res. 2012, 45, 2211–21.

[9] R. L. E. Furlan, S. Otto, J. K. M. Sanders, Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 4801–4. [10] S. Otto, R. L. E. Furlan, J. K. M. Sanders, Science 2002, 297, 590–593.

[11] A. R. Lippert, A. Naganawa, V. L. Keleshian, J. W. Bode, J. Am. Chem. Soc. 2010, 132, 15790–15799. [12] C. S. Mahon, D. A. Fulton, Chem. Sci. 2013, 4, 3661–3666.

[13] S. Hamieh, V. Saggiomo, P. Nowak, E. Mattia, R. F. Ludlow, S. Otto, Angew. Chem. Int. Ed. 2013, 52, 12368–12372. [14] F. Ulatowski, A. Sadowska-Kuziol̷a, J. Jurczak, J. Org. Chem. 2014, 79, 9762–9770.

[15] I. A. Riddell, T. K. Ronson, J. K. Clegg, C. S. Wood, R. A. Bilbeisi, J. R. Nitschke, J. Am. Chem. Soc. 2014, 136, 9491– 9498.

[16] L. Ratjen, G. Vantomme, J.-M. Lehn, Chem. Eur. J. 2015, 21, 10070–10081. [17] R.-C. Brachvogel, F. Hampel, M. von Delius, Nat. Commun. 2015, 6, 7129.

[18] D. J. Hansen, I. Manuguerra, M. B. Kjelstrup, K. V. Gothelf, Angew. Chem. Int. Ed. 2014, 53, 14415–14418. [19] M. Mondal, A. K. H. Hirsch, Chem. Soc. Rev. 2015, 44, 2455–2488.

[20] Z. Grote, R. Scopelliti, K. Severin, Angew. Chem. Int. Ed. 2003, 42, 3821–3825. [21] K. Severin, Chem. Eur. J. 2004, 10, 2565–2580.

[22] I. Saur, K. Severin, Chem. Commun. 2005, 1471–1473.

[23] P. T. Corbett, S. Otto, J. K. M. Sanders, Chem. Eur. J. 2004, 10, 3139–3143. [24] P. T. Corbett, J. K. M. Sanders, S. Otto, J. Am. Chem. Soc. 2005, 127, 9390–9392. [25] P. T. Corbett, J. K. M. Sanders, S. Otto, Chem. Eur. J. 2008, 14, 2153–2166. [26] J. Solà, M. Lafuente, J. Atcher, I. Alfonso, Chem. Commun. 2014, 50, 4564–4566.

[27] M.-K. Chung, P. S. White, S. J. Lee, M. R. Gagné, Angew. Chem. Int. Ed.) 2009, 48, 8683–8686.

[28] F. B. L. Cougnon, H. Y. Au-Yeung, G. D. Pantoş, J. K. M. Sanders, J. Am. Chem. Soc. 2011, 133, 3198–3207. [29] J.-F. Ayme, J. E. Beves, D. A. Leigh, R. T. McBurney, K. Rissanen, D. Schultz, Nat. Chem. 2012, 4, 15–20. [30] N. Ponnuswamy, F. B. L. Cougnon, J. M. Clough, G. D. Pantoş, J. K. M. Sanders, Science 2012, 338, 783–785. [31] N. Ponnuswamy, F. B. L. Cougnon, G. D. Pantoş, J. K. M. Sanders, J. Am. Chem. Soc. 2014, 136, 8243–8251. [32] J. W. Sadownik, D. Philp, Angew. Chem. Int. Ed. 2008, 47, 9965–9970.

[33] B. Rubinov, N. Wagner, H. Rapaport, G. Ashkenasy, Angew. Chem. Int. Ed. 2009, 48, 6683–6686.

Referenties

GERELATEERDE DOCUMENTEN

To uncover the selection criteria for the emergence of peptide-based replicators in dynamic combinatorial libraries, we have extended our portfolio with thirteen novel peptides

In the system, which is based on a tyrosine-based peptide building block, emergence of the replicator has no spe- cific dependence on the structure of the other replicator but

In the system, which is based on a tyrosine-based peptide building block, emergence of the replicator has no spe- cific dependence on the structure of the other replicator but

In tegenstelling tot deze zeer specifieke interactie tussen replicatoren, hebben we in Hoofdstuk 4 laten zien hoe de grootte van nieuwe replicatoren kan worden bepaald door de

The number of citations that a scientist receives is a better quality indica- tor than the number of articles (s)he has published.. However, neither fully reflect the quality of

Self-replicators from dynamic molecular networks: selection, competition and subsystem coupling.. University

This introduction gives a short overview about the most important types of dynamic combinatorial chemistry (as the covalent processes enabling reaction networks) and of

Here we report that coupling two dynamic combinatorial subsystems, featuring two separate building blocks, enables effector-mediated control over the onset of self-replication..