+ All Categories
Home > Documents > Green Solvents in Peptide Synthesis

Green Solvents in Peptide Synthesis

Date post: 19-Jan-2022
Category:
Upload: others
View: 11 times
Download: 0 times
Share this document with a friend
46
Green Solvents in Peptide Synthesis Lillian Worst Submitted in Partial Fulfillment of the Prerequisite for Honors in Biochemistry under the advisement of Julia Miwa May 2021 © 2021 Lillian Worst
Transcript

Green Solvents in Peptide Synthesis

Lillian Worst

Submitted in Partial Fulfillment of the

Prerequisite for Honors in Biochemistry

under the advisement of Julia Miwa

May 2021

© 2021 Lillian Worst

1

COVID-19 impact statement

Due to the COVID-19 pandemic, the scope of this project had to be scaled back. Initially we planned on obtaining quantitative data on the use of propylene carbonate in peptide synthesis. However, due to my late start to this thesis (I originally started a thesis in a different lab, but had to switch due to the pandemic), a combined 3 quarantines between Professor Miwa and myself, and impacted supply chains for several of the required reagents, we were unable to begin collecting data until late March 2021. Though it is not the project we initially envisioned, it still lays a good groundwork for future research into propylene carbonate and its use as a green solvent in peptide synthesis.

2

Abstract Peptide synthesis is one of the most widely used and important sets of organic

reactions in biochemistry. With demand for peptides increasing, it is imperative that peptide synthesis becomes a greener process, better for both humans and the environment. One major means of “greening” peptide synthesis involves replacing the toxic solvent dimethylformamide (DMF) with a greener alternative. Previously, propylene carbonate, a very green solvent, has shown promise as a replacement for DMF in solution-phase peptide synthesis. This thesis focuses on the use of propylene carbonate as a solvent in the more prevalent solid-phase peptide synthesis, as well as the adaption of a substitution level assay to determine the quantitative success of solid-phase coupling in propylene carbonate. It was determined that solid-phase synthesis in propylene carbonate is possible, and that the substitution level assay can be adapted to work in propylene carbonate. Furthermore, all necessary reagents for peptide synthesis were shown to dissolve in propylene carbonate, making it a strong contender to replace DMF as the leading solvent for peptide synthesis. Future work will determine the quantitative comparison between propylene carbonate and DMF as solvents, and adapt the solid-phase peptide synthesis reactions for improved yields in propylene carbonate.

3

Table of Contents

Introduction 5 Peptides 5 Principles of Green Chemistry 6 Peptide Synthesis 9 Solvents in Peptide Synthesis 15 Coupling 17 Capping and Deprotection 23 Monitoring Peptide Synthesis 26 Specific Goals of the Thesis 27

Results and Discussion 28

Future Directions 36

Experimental Procedures 38 Coupling of amino acid to Chemmatrix Resin (general procedure) 38 Deprotection of Fmoc group from Fmoc-aminoacyl-chemmatrix resin: 39 Solubility Tests 39 NMR 40

References 41

4

Acknowledgements

I have many people I would like to thank for their help in this process. First, I would like to thank Professor Miwa for being an amazing teacher and mentor these last four years. Thank you for being a wonderful professor and agreeing to restart your lab to work on a thesis with me in these unprecedented times. I would also like to thank the members of my committee, Adam Matthews and Elizabeth Oakes, for meeting with us and helping us to get the project moving along. You are both wonderful professors and I am honored that you were willing to be a part of this project. Thank you also to Professors Chang and Radhakrishan for agreeing to attend my defense.

I need to thank my family and friends as well. Thank you to my parents and brother for supporting me over the phone and putting up with all of my nonsense for the last 22 years. I am sorry that I had to change projects just at the point where you were starting to understand what my first one was about, I hope you like this one too. Thank you to my friends, especially Ally, Gabe, and Landon, who have been amazing to block with through this pandemic and have been incredible sounding boards for some of my wilder ideas these past few months. I am looking forward to starting the next chapter of our lives together.

5

Introduction

Peptides

Peptides are short chains of amino acids, the building blocks for proteins. The

difference between peptides and proteins is fuzzy at best, but it is often determined by

length, manufacture, and use. Peptides are used commonly in many types of research

including helping to understand protein domains and epitopes (Tanabe, 2007), create

the next generation of antibiotics (Zhang, 2016), design new enzymes and binding sites

(Zhang et al., 2014), and activate immune responses (Tanabe, 2007) among other

things (PBRL). Peptides are likely to play an increasingly important role in medicine,

with 60 peptides currently on the market as therapeutics for a variety of conditions from

HIV to cancer, and over 600 more in development or clinical trials (Fosgerau et al.,

2015). Peptides make for great therapeutics due to their extreme specificity for one

target, which minimizes off-target interactions and side effects. Despite the promise of

peptides as medication, they are only just now beginning to hit the market due to the

difficulty of delivery to cells when taken orally. Enzymes in our stomachs are specifically

designed to break down peptides, meaning that strategies such as peptide stapling and

use of unnatural amino acids had to be developed before they could become viable as a

class of medicines (Cromm et al., 2015).

Both proteins (≥ 50 amino acids) and peptides (< 50 amino acids) are readily

prepared in the research laboratory. While proteins are most commonly produced by

cloning in bacterial cells, peptides are primarily produced via organic synthesis. While

peptide and protein production in cells is efficient and fast, peptide synthesis is much

more versatile, allowing for the addition of unnatural amino acids and modifications to

6

the peptide backbone. As processes such as peptide stapling rely on backbone

modifications and incorporation of unnatural amino acids, peptide synthesis is

increasingly important in research. Many academic and industrial labs complete peptide

synthesis on a large scale. This means that peptide synthesis is a common and

necessary set of organic reactions that is widely used. Any improvement to the process

of peptide synthesis has the potential to have a widespread impact.

Principles of Green Chemistry

Organic chemistry has been a distinct discipline of chemistry since the 1800s, but

only in the last 30 years has green chemistry become a concern to organic chemists

(Anastas and Beach, 2017). Green Chemistry was originally a response to the passage

of the Pollution Prevention Act of 1990, which aimed to improve the design of industrial

processes to decrease their toxicity and harm to human health and the environment.

This new effort contrasted with previous policies on pollution prevention that aimed to

mitigate the effects of harmful substances after the fact instead of designing the

processes themselves to be less harmful (Anastas and Beach, 2017). Throughout the

1990s, interest and funding for green chemistry grew, and in 1998, the 12 Principles of

Green Chemistry (Figure 1) were published as a guideline for future research. Overall,

the 12 principles provide criteria for designing new reactions and improving existing

reactions to decrease their environmental impact and potential harm to humans. There

are several different approaches to the goals that are laid out in the 12 principles. For

example, the first two principles, reducing waste and improving atom economy, both

aim to reduce the quantity of byproducts of a reaction. Waste prevention focuses more

7

on reducing the use of solvents and auxiliaries, while atom economy advocates for the

creation of reactions that do not produce unwanted byproducts, meaning that 100% of

the atoms present as reactants can be found in the products. Both of these principles

aid in creating reactions that are more efficient and generate less hazardous waste than

traditional syntheses. Efficient reactions also produce fewer byproducts. Byproducts and

reactants can be hard to separate from the product, requiring larger amounts of solvent

and other resources that become waste. Decreasing the number of byproducts

produced is therefore green in several ways. First, it reduces the waste of the

byproducts themselves, and second it decreases the solvent necessary in the reaction

and subsequent work-up, which would have otherwise become waste. Therefore, a

8

more efficient reaction that does a better job of converting the reactant into the desired

product will use fewer resources and produce less waste.

Other principles focus more on decreasing the toxicity of all organic species that

might be used during the reaction, from solvents, to waste, to byproducts, and even

products. One way to reduce the toxicities of these species is to find less toxic

alternatives to staples used in the reaction. Any alternative will need to have similar

properties to the original compound, but a lower toxicity. One example of this comes

from the field of solar cell development. Many of the compounds necessary for solar

cells tend to be both toxic and costly (defeating some of the purposes of solar energy).

Much chemical research in the last few decades has focused on the semiconductors

used in solar cells. Recently, Lokhande et al. have described replacing toxic CdTe and

CuInGaS2 (CIGS) quaternary chalcogenide compounds with the much safer Cu2ZnSnS4

compounds as semiconductors, therefore creating a much safer and greener production

process (Lokhande et al., 2016). This example also demonstrates another principle of

green chemistry: that if a new reaction is being designed from scratch, a good way to

decrease toxicity is simply to design reactions that do not require toxic materials. In the

case of Cu2ZnSnS4, it was chosen for use in solar cells partially due to the relative

abundance of its components, and relative ease of synthesis (Lokhande et al., 2016).

Unfortunately, designing reactions for a decrease in toxicity can lead to a

decrease in yield, which may increase waste. In other words, the principles of green

chemistry that call for a decrease in toxicity of products can directly conflict with the

principles that call for a decrease in waste, which is an important consideration

whenever trying to develop a greener reaction. However, using fewer toxic materials

9

ensures that even if more waste is produced, it will be less harmful to humans and the

environment, and usually easier to recycle or dispose of. Furthermore, once reactions

are initially developed for use of less toxic materials, yields can be incrementally

improved in the same way all other chemical reactions have been improved by

perfecting the reaction conditions over time.

Other principles of green chemistry consider having renewable feedstocks and

reducing the energy required to carry out a reaction by decreasing the heat or pressure

or using renewable sources of electricity. Overall, the principles of green chemistry

provide metrics for improving reactions to reduce their negative impacts on human

health and the environment. Greener chemical reactions are truly the future of

chemistry. It is vital that reactions developed before the emergence of green chemistry

be improved and made safer for humans and the environment. Since peptide synthesis

is such a widely used and important set of chemical reactions, it provides a great target

for improvement using green chemistry principles.

Peptide Synthesis

Both peptide synthesis and the biological production of proteins and peptides

require joining the carboxyl group of one amino acid to the amino group on another,

creating an amide (also referred to as a peptide) bond (Figure 2). Despite this

overarching similarity, the reactions carried out in peptide synthesis differ from those

carried out by cells in two key ways. First, chemical synthesis proceeds from the C-

terminus to the N-terminus while ribosomal translation occurs from the N-terminus to the

C-terminus. Second, because any chemical catalysts or coupling reagents used in

10

peptide synthesis cannot achieve the same amount of specificity as an enzyme, amino

acids for peptide synthesis must have protecting groups covering all reactive functional

groups except the one currently reacting to form the peptide bond. Specifically, the

carboxyl group on one amino acid, and the amino group on the other need to be

protected as well as any side chain functional groups that might react (such as the

carboxyl group on glutamic acid). These protecting groups ensure that only the correct

amino acids react with each other during the coupling process. Since peptides are

synthesized from C to N, the C-terminal protecting group must remain in place for the

entirety of the synthesis process (as do side chain protecting groups), while the N-

terminal protecting group must be removed after each step so that they next amino-

protected amino acid can be added. Thus, the synthesis requires orthogonal protection,

in which the “temporary” amino protecting group can be removed without removing the

“semi-permanent” protecting groups on the side chains and C-terminus. Therefore, even

though both biological synthesis of proteins and peptides essentially repeats the same

coupling step over and over, peptide synthesis also requires a deprotection step

between each coupling step.

Protecting groups affect the solubility of some amino acids. Since protecting

groups are often hydrophobic and they cover a hydrophilic amino group on each of the

11

incoming amino acids, protected amino acids are much less soluble than their

unprotected counterparts in a variety of solvents, including those commonly used in

peptide synthesis. This restricts the solvents that can be used in peptide synthesis,

especially when using the more common Fmoc-protected amino acids, which are

especially hard to dissolve in most solvents.

One important constraint of peptide synthesis is that it requires very high yield at

every step. This requirement is due to the sheer number of steps necessary to create

some peptides. While peptides are traditionally defined as under 50 amino acids, a

larger peptide could spell over 100 steps in a synthesis. A 50 amino acid peptide with a

95% yield of each coupling step could have an overall yield of less than 8%. In addition,

incomplete coupling steps lead to deletion impurities, meaning that the resulting peptide

will only be missing a single amino acid in the middle, which can be extremely difficult to

separate from the desired end product. In order to prevent the inclusion of these

deletion impurities, long synthesis processes often include a capping step after each

coupling step, wherein the unreacted free amine is capped with a third type of protecting

group. Capping changes deletion impurities to truncation impurities, making them much

easier to separate from the final product, but does not help maintain high yield. The

necessity of extremely high yield makes peptide synthesis a more difficult target for

green chemistry, as it is often the case that any change to a synthetic process will

initially decrease its yield. Since changes to any chemical process almost always initially

result in a decrease in the yield, all new techniques must be perfected in order to be

useful on an industrial scale.

12

Peptide synthesis was revolutionized in 1963 when R. B. Merrifield invented

solid-phase peptide synthesis (SPPS). Solid-phase peptide synthesis makes industrial

scale peptide synthesis possible by simplifying the purification processes and therefore

increasing the overall yield. Solid-phase synthesis involves anchoring the C-terminal

amino acid of the peptide to an insoluble resin bead. The first resin used for SPPS was

a functionalized polystyrene, and almost all SPPS resins in use today are derivatives of

polystyrene. The attachment of the peptide to the insoluble resin allows for easy

separation of the growing peptide chain from all solvents, unused reagents, and

byproducts. At each step, the solid resin with the peptides attached is simply filtered out

of the solution. Solid phase peptide synthesis can also increase yield by use of an

excess of reagents, forcing the reaction to completion through LeChatelier's principle.

Like other forms of peptide synthesis, SPPS involves repeated coupling and

deprotection steps (Figure 3).

Even though the chemistry of solid-phase peptide synthesis is essentially the

same as solution-phase peptide synthesis, solid-phase synthesis has several unique

concerns. First, the resin has to swell in the presence of the solvent, allowing for the

reagents to enter the interior of the resin matrix and react with the peptides anchored

there (as well as the ones on the surface) (AAAPTECH). The polystyrene resins used

in SPPS swell to large volumes in dichloromethane and dimethylformamide, solvents

commonly used in SPPS. In contrast, the resin shrinks to a small volume in methanol,

making it an unsuitable solvent for SPPS reactions.

Next, since SPPS relies on being able to separate the solid resin from the liquid

solution based on phase, it is imperative that no solid side products are produced during

13

any point in the synthesis (El-Faham and Albericio, 2011). Even with these constraints,

solid-phase peptide synthesis is the far superior form of peptide synthesis in most

cases, due to its increased yields and ease of application. Despite the fact that solution-

phase peptide synthesis was developed first, solid-phase peptide synthesis is the

industry (and academic) standard. Thus, applying green chemistry to solid-phase

peptide synthesis will have much more of an impact than applying it to solution-phase

peptide synthesis.

14

In 2017 Lawrenson et al. produced a paper focused on using cyclic carbonates, a

type of green solvent, in peptide synthesis (Lawrenson et al., 2017). Currently, the most

commonly used solvent for peptide synthesis is dimethylformamide (DMF) which is both

extremely toxic and hard to recycle. DMF is the leading standard solvent in peptide

synthesis because it easily dissolves Fmoc-protected amino acids (which are less

soluble in other organic solvents). Due to the large amounts of DMF used during the

peptide synthesis process, replacing it with a greener solvent could have a significant

15

impact on the overall greenness of the reaction. However, though Lawrenson et al. did

describe an environmentally friendly solvent that can replace DMF, they focused mainly

on solution-phase peptide synthesis. Though this was a good initial proof of concept, it

is imperative that cyclic carbonates and other green solvents be tested in solid-phase

peptide synthesis, since it is the predominant way peptides are manufactured in both

industry and academia.

Solvents in Peptide Synthesis

Replacing a solvent in an established reaction requires a knowledge of the

specific qualities of the original solvent. As previously mentioned, the most commonly

used solvent in peptide synthesis is currently dimethylformamide (DMF), though

dichloromethane is also used at certain points in the process. Both DMF and

dichloromethane are aprotic, polar solvents. This means that they have the ability to

dissolve a wide range of molecules, including both hydrophilic and hydrophobic

molecules, but they lack the ability to donate a proton to their solutes (Atherton).

Because of the protecting groups on the amino acids, this ability to dissolve a wide

range of hydrophobic and hydrophilic molecules is critical to the success of a peptide

synthesis reaction. Fmoc-protected amino acids especially are difficult to dissolve in

many polar solvents due to the hydrophobicity of the Fmoc group. On the other hand,

many coupling reagents are salts, meaning that they struggle to dissolve in nonpolar

solvents. DMF (and other polar aprotic solvents) tend to do a good job of spanning

these differences and dissolving both the hydrophobic protected amino acids, and the

hydrophilic coupling reagents. Thus, any solvent used in place of DMF must be similarly

16

polar and aprotic. It would also be beneficial to match other properties of the solvent,

such as the melting and boiling points. This is because the methods used for peptide

synthesis in DMF have been perfected for years, down to temperature, run time, and

concentrations. Deviation may at least temporarily decrease yield, which may

discourage the widespread adoption of greener solvents.

One family of candidates that has emerged as a green alternative to DMF in

peptide synthesis is cyclic carbonates (Figure 4). Cyclic carbonates are extremely green

solvents due to their relative safety, ease of recycling, and 100% atom economical

synthesis that sequesters CO2 (Figure 5). In fact, propylene carbonate was ranked as

the greenest solvent in 2014 (Prat et al., 2014). In 2017, Lawrenson et al. produced a

paper describing the use of propylene and ethylene carbonate as substitute solvents in

peptide synthesis (Lawrenson et al., 2017). Their initial findings suggest that propylene

carbonate holds great potential to replace DMF as a solvent in peptide synthesis, as

long as the process can be modified. Many of their qualitative results for synthesis in

propylene carbonate were comparable to synthesis in DMF (Lawrenson et al., 2017).

However, their work focused mostly on comparing propylene carbonate to another

17

cyclic carbonate, ethylene carbonate, in solution phase synthesis. They did gather some

data that shows solid-phase synthesis can be carried out in propylene carbonate but did

not quantitatively show its effects on the synthesis when compared to DMF or use

modern coupling reagents. Since solid-phase peptide synthesis is more commonly used

both industrially and in academic settings to make peptides, it is important to test this

solvent more quantitatively using common solid-phase reactions and coupling reagents.

Coupling

The key step in any peptide synthesis is the coupling of the two amino acids to

one another. In cells, this step is carried out by the ribosome, but in peptide synthesis, it

is facilitated by chemical catalysts known as coupling reagents. Coupling reagents work

by activating the carboxyl group of the incoming amino acid so that it can form the

peptide bond with the amino group of the growing peptide. On top of this, the coupling

step involves formation and breakage of bonds near the alpha carbon of the amino acid.

This means that for some amino acids such as histidine, there is a good chance that

side reactions can occur with the alpha carbon leading to racemization of the amino

acid (El-faham and Albericio, 2011). This can be a huge problem since peptides, like all

proteins, are stereospecific and derive their function from their three-dimensional

18

structure. Racemization of the amino acids during the coupling reaction could lead to a

significant decrease (up to 50%) in the yield for each step. Thus, coupling reagents that

have lower rates of racemization are preferred in modern peptide synthesis. There have

been many generations of coupling reagents, with the most important shown below

(Figure 6). The Lawrenson et al. paper from 2017 primarily relied on the use of ethyl-3-

(3-dimethylaminopropyl)carbodiimide (EDC) and dicyclohexylcarbodiimide (DCC) along

with various coupling additives (Figure 6). The carbodiimides DCC and EDC were two

of the original coupling reagents, with use in peptide synthesis dating back to the 1950s

(El-faham and Albericio, 2011). However, carbodiimides, and DCC in particular, are not

the best reagents available for solid-phase peptide synthesis. Their higher rates of

racemization can lead to a decrease in yield overall. Carbodiimide coupling occurs

through an addition of a carbonyl to the carboxyl oxygen of the amino acid and can

have unwanted side products that further decrease yield (Figure 7). While coupling

additives (Figure 6) can help decrease racemization, newer coupling reagents are less

complicated to use and often produce higher yields, making them preferable in most

cases. DCC is particularly unsuited to solid phase peptide synthesis, as one of the

natural side products of coupling (dicyclohexylurea) is soluble only in trifluoroacetic acid.

This means that in the absence of large amounts of trifluoroacetic acid (which is often

only used as a final step in synthesis to remove orthogonal protecting groups), it forms a

19

solid that cannot be separated from the resin by filtration (El-Faham and Albericio,

2011).

The use of these older coupling reagents may help to explain the lower yields

(between 75 and 80%) in the results from Lawrenson et al, and replacing the coupling

20

reagent may help improve the yield for propylene carbonate. Furthermore, it is important

to see if propylene carbonate performs similarly to DMF under conditions that would be

more commonly practiced during peptide synthesis in the lab or in an industrial setting,

as many of the techniques for use of coupling reagents have been perfected in DMF.

21

From a review of the literature, we determined that two of the most commonly used

coupling reagents in modern solid phase peptide synthesis are benzotriazol-1-yloxytris-

pyrrolidino-phosphonium hexafluorophosphate (PyBOP), a phosphonium salt, and

tetramethyluronium hexafluorophosphate (HBTU), a uronium salt (Figure 8A, Figure 6).

Both have similar mechanisms for activating the carboxyl group on the amino acid

(Figure 8B-C). The charged nitrogen on the coupling reagent facilitates the formation of

a temporary covalent bond with the carboxyl group on the amino acid. The carboxyl

group is then activated through the transfer of the nitrogen ring system to the carboxyl

oxygen. In both cases, the intermediate ester is also activated, and able to form a

peptide bond. However, due to the short lifespan of these intermediates, the peptide

bond formation is much more likely to occur with species 17 which is labeled as the

activated leaving group. Both HBTU and PyBOP have much lower rates of

racemization than carbodiimides due to the fact that they contain the racemization

suppressant HOBt (Figure 6) in their structure. The first coupling reagent to employ this

trick was BOP, the direct predecessor of PyBOP. Although BOP was a fantastic

coupling reagent, it produced a carcinogenic byproduct (hexamethylphosphoramide)

where PyBOP produces the relatively harmless species 19 (El-Faham and Albericio,

2011). Thus, PyBOP and several other similar phosphonium reagents were developed

to eliminate the production of toxic byproducts. This is an excellent example of ways in

which peptide synthesis has already been modified to make it a greener process. Unlike

carbodiimides, both PyBOP and HBTU have no side reactions that decrease the overall

yield of the reaction. Furthermore, they both have low racemization rates and are

relatively inexpensive as coupling reagents go, meaning that they are widely used for

22

peptide synthesis (Palasek et al., 2006 and Coste et al., 1990). Both PyBOP and HBTU

must prove effective in propylene carbonate in order for it to be a viable replacement for

DMF.

Another constraint of all coupling reagents is that not all coupling reactions are

created equally. Some pairs of amino acids are specifically more difficult to couple than

others due to the structures of their side chains, and their position in the growing

23

peptide. For example, it has been widely observed that using carbodiimide coupling

reagents leads to a difficulty in completing additions of amino acids 12-20 to the growing

peptide (Dunn et al., 1994). Furthermore, hydrogen bonding between unprotected side

chains such as those found on Gln and Asn can also lead to a decrease in the success

of an individual coupling step due to the formation of secondary structures (Dunn et al.,

1994). Many techniques are employed to help improve the success of each individual

coupling step, including using different coupling reagents for specific steps. Another

advantage of HBTU and PyBOP is their general fitness to most coupling steps. Though

some other coupling reagents may be better suited to a specific difficult coupling step,

HBTU and PyBOP (along with other phosphonium and uronium reagents) have the

ability to do a relatively good job coupling most steps (El-Fahan and Albericio, 2011).

This general promiscuity, combined with their low rates of racemization, relatively low

cost, and lack of toxicity (both themselves and in their by-products) have ensured that

they are widely used coupling reagents by most people doing peptide synthesis. Thus,

any green protocol for peptide synthesis must be compatible with both HBTU and

PyBOP.

Capping and Deprotection

The other major step in peptide synthesis is deprotection. Deprotection involves

removing the N-terminal protecting group from the growing peptide after each coupling

step, allowing for the next coupling to take place. There are two main types of protection

strategies commonly used in peptide synthesis: Boc/Benzyl and Fmoc/t-butyl. In both

cases, the first protecting group listed (Boc, Fmoc) refers to the temporary N-terminal

24

protecting group, and the second (benzyl, t-butyl) refers to the orthogonal semi-

permanent protecting groups on the side chains. In each case, the two types of

protecting groups are removed through different processes, allowing the N-terminal

protecting group to be removed before each coupling step while leaving the orthogonal

protection to the end (Albericio, 2000). Boc/benzyl protection had mostly fallen out of

favor for use in solid phase peptide synthesis due to the necessity of harsh conditions

required for deprotection of the amino group (trifluoroacetic acid), the necessity to

neutralize the acid before coupling can occur (Figure 9), and the requirement of a very

dangerous hydrofluoric acid procedure for the final removal of the side chain protecting

groups and cleavage of the peptide from the resin. Fmoc/t-butyl protection only requires

harsh conditions (trifluoroacetic acid) for the final cleavage from the resin and

deprotection of the orthogonal t-butyl groups. Since the Fmoc group is removed in

piperidine (a base), while the resin and side chain protecting groups are removed with

acid, the protection strategy is truly orthogonal. Furthermore, the basic conditions

necessary for coupling with reagents such as PyBOP are already present, and there is

no need to first neutralize the solution (Figure 9). Unfortunately, Fmoc protected amino

acids are harder to dissolve in most solvents when compared with Boc protected amino

acids. This led to the widespread adoption of DMF in peptide synthesis, as it is one of

the few solvents that dissolve Fmoc protected amino acids extremely well. Thus, it is

imperative that propylene carbonate demonstrate the ability to dissolve Fmoc amino

acids as well.

25

Another important step in the peptide synthesis process is capping. Capping

involves blocking the end of an unreacted peptide so that it cannot continue to react.

This prevents deletion impurities, which are much harder to separate from the final

product than the much shorter truncated impurities. Capping is usually accomplished by

addition of acetic anhydride, which reacts with the free amine to create an acetamide

group that cannot be removed by deprotection steps, thus truncating the peptide

(Verma et al., 2020). Capping is most important in the synthesis of long peptides where

deletion impurities would be much harder to separate from the desired product.

26

Ultimately, solid-phase peptide synthesis is a complex scheme of chemical

reactions that result in the ability to create any peptides regardless of sequence and

easily separate them from waste and byproducts. It is a powerful tool, and one that will

become increasingly important with the adoption of peptide medicines, but currently has

significant negative effects on people and the environment. It is imperative that greener

methods of peptide synthesis are developed and implemented.

Monitoring Peptide Synthesis

One of the most important parts of any complex reaction involves the ability to

monitor its progress and quantify its success after each stage. There are several

common methods for monitoring peptide synthesis reactions. The Kaiser test is a

commonly used method to monitor peptide synthesis. After each coupling step,

ninhydrin is added, which turns blue in the presence of free amine. By either looking at

the color of the solution or taking a spectrum, the general success of the coupling step

can be determined. Though the Kaiser test is very fast, and extremely useful

qualitatively, it does not provide quantitative results on the success of a coupling

reaction and can be inaccurate when used for certain amino acids (Fontenot et al.,

1997). On the other end of the spectrum, high performance liquid chromatography

(HPLC) can be used to get good quantitative data on the success of a coupling reaction

but is also very time and labor expensive. A happy medium can be struck by using a

substitution level analysis, which uses the fact that the byproduct of Fmoc deprotection,

dibenzofulvene, absorbs UV light. By using spectroscopy on an aliquot of deprotection

solution removed from the reaction vessel, it is possible to get some quantitative data

27

on the success of a peptide coupling reaction in much less time than it would take to

perform HPLC.

Specific Goals of the Thesis

This thesis will primarily focus on the methods needed to obtain quantitative data

from solid phase peptide synthesis in propylene carbonate. It will assess the solvation of

HBTU, PyBOP, Fmoc amino acids, and piperidine in propylene carbonate, as well as

address the use of ChemMatrix rink amide resin (which was selected due to its

compatibility with propylene carbonate). Furthermore, we will attempt to develop and

improve on a substitution level assay necessary to determine the success of a coupling

in propylene carbonate. Ultimately, we hope to prove that propylene carbonate can be a

viable replacement for DMF in modern solid phase peptide synthesis and set the stage

for a quantitative study directly comparing propylene carbonate to DMF.

28

Results and Discussion

The goal of this project was to adapt a solid-phase peptide synthesis protocol

using modern coupling reagents and the green solvent propylene carbonate. In order to

succeed in propylene carbonate, a solid phase synthesis reaction has several specific

requirements. First, a solid support resin must be able to demonstrate the appropriate

swelling in propylene carbonate. For this project, we used H-Rink Amide ChemMatrix

resin (same as Lawrenson et al).. ChemMatrix is a polyethylene glycol (PEG) based

resin rather than a polystyrene-based resin. It demonstrates excellent swelling in

propylene carbonate. Second, all reagents (protected amino acids and coupling

reagents) had to be able to dissolve in propylene carbonate at useful concentrations for

peptide synthesis. Third, the temporary N-terminal protecting groups (Fmoc in this case)

must be removable in propylene carbonate. Finally, there must be a reliable method to

measure the extent of coupling (the substitution level of protected amino acid on the

resin). The work done during this thesis project has focused primarily on the

development of the assay, assessment of solubility of a variety of protected amino

acids, and demonstration of successful deprotection and coupling on a sample synthetic

peptide.

Initially, we planned to only test the solubility of the reagents we planned to

actually use in the final coupling reaction. This included Fmoc-Val-OH, Fmoc-Phe,

PyBOP, HBTU, and tygon tubing (which we were hoping did not dissolve). Though the

results of this first round of solubility tests indicated coupling reagents and Fmoc-amino

acids can be dissolved in propylene carbonate, it also indicated that not all Fmoc amino

acids dissolve in propylene carbonate equally well. Specifically, in order to get the

29

Fmoc-Phe to dissolve, it was necessary to add a small amount of DMF. Thus, another

round of solubility tests was carried out, this time using all of the Fmoc amino acids on

hand (13 of the 20 naturally occurring amino acids were tested). To conserve solvent,

the concentration of amino acid typically present in a coupling reaction was calculated

(0.1M), and the amount to dissolve in 1 mL of propylene carbonate was determined

(0.0001 mol). The amino acid (0.0001 mol) was added to 1.0 mL of propylene

carbonate. Each sample was sonicated for one minute in an ultrasonic bath. Afterwards,

the samples that had dissolved were noted. For the samples that had not dissolved,

another 1 mL of propylene carbonate was added, and the samples were again placed in

the ultrasonic bath. Again, it was noted which amino acids had dissolved. For the five

amino acids that remained undissolved, DMF was added dropwise until either the amino

acid dissolved, or 100 drops was reached (Table 1). For arginine (Fmoc-Arg), it took

over 100 drops of DMF and would still not dissolve. An equivalent amount of the amino

acid was again weighed out and this time dissolved in 1mL of pure DMF to ensure that it

would dissolve in DMF. It did successfully dissolve in DMF, and subsequent dropwise

addition of 100 drops of propylene carbonate did not cause it to crash out of the

solution, suggesting that propylene carbonate could still be used to dilute the DMF, and

therefore make the reaction greener. It does seem that the propylene carbonate has

trouble dissolving any species with an unprotected nitrogen based functional group, as

the other amino acid that needed a significant amount of DMF to dissolve was

asparagine. These results are still very good however, as it is clear that PC can dissolve

the majority of the Fmoc-protected amino acids with minimal addition of DMF, and there

is a good chance that a greener cosolvent can be discovered in the future.

30

The quantitative assay chosen to measure the amount of protected amino acid

(in mmol amino acid per gram of resin) involves taking a specific mass of resin,

deprotecting the Fmoc group with piperidine, diluting the supernatant and using the UV

absorbance of the dibenzofulvene produced to quantify the amount of Fmoc groups that

were removed from the resin. This assay had been used successfully in the Miwa lab

with Fmoc-aminoacylated polystyrene resins that were deprotected with piperidine in

DMF. The primary focus of the project was to adapt this substitution level assay using

propylene carbonate and ChemMatrix resin. The substitution assay aimed to

quantitatively determine the success of a coupling step and was evaluated using two

31

different resins: Fmoc-Phe-polystyrene (a resin commonly used in traditional peptide

synthesis) and Fmoc-Phe ChemMatrix (a resin that was chosen due to its compatibility

with propylene carbonate synthesis), which was prepared in lab from a commercially

available H-Rink amide ChemMatrix resin. Fmoc-Phe-polystyrene is a commercially

available resin, so there was an ample supply of this resin, all with the same (known)

substitution level of 0.76 mmol/g resin. Because the Fmoc-Phe ChemMatrix resin was

prepared in the lab, the substitution level was unknown. However, the H-Rink amide

ChemMatrix resin from which it was prepared had a substitution level of 0.4 mmol/g

resin, meaning that in direct comparisons between the H-Rink amide ChemMatrix resin

32

and the Fmoc-Phe-polystyrene resin, we would expect the ChemMatrix resin to have a

lower substitution level. While the two resins had very similar spectra, ultimately, the

Fmoc-Phe-Polystyrene in DMF gave more clustered data, while the ChemMatrix resin in

PC had more variation (Figure 10). This was at least partially due to the difference in

size of the resin particles. While the Fmoc-Phe-Polystyrene had a very small particle

size and tended to settle to the bottom in the DMF, the ChemMatrix resin had a much

larger particle size, especially when swelled, and remained suspended in the PC. This

led to difficulty in accurately pipetting aliquots of deprotection solution without plugging

the tip of the micropipette, or accidentally incorporating resin. An additional filtration step

after deprotection did decrease the variability of the spectra somewhat, though it was

still not as consistent as the data taken in DMF.

Initially, the substitution level assay had wide variation using Fmoc-Phe-

polystyrene in DMF. In order to improve the reproducibility and precision of the assay,

two major changes were made. The first goal was to improve the weighing, transfer,

and handling of the resin. Both the polystyrene resin and the ChemMatrix resin are

extremely likely to stick to the spatulas and the sides of the test tubes. This impacted

the accuracy of the measured masses of resin in each test tube, as resin on the side of

the test tube contributed to the overall mass of the resin in the tube but was unable to

be washed down into the reaction solution, and thus did not contribute to the amount of

dibenzofulvene in the end solution. In order to fix this problem, the following changes

were made. First, instead of diluting only one aliquot from each 1.00 mL sample, the

volume of deprotection solution was increased to 2.00 mL, and multiple (4-5) aliquots

were diluted and tested from each sample. This change ensured that any error resulting

33

from resin sticking to the test tube would be systematic rather than random. Second, the

original borosilicate test tubes were switched for microcentrifuge tubes with a much

smaller volume and therefore smaller surface area on the sides for the resin to stick.

The combination of these two changes helped decrease the noise in the data.

Further improvement was made by replacing the old piperidine with fresh

piperidine. Time trial experiments carried out with old piperidine indicated that 20

minutes may not have been adequate to complete deprotection, as samples left

overnight showed significant improvement over those that were only allowed to

deprotect for 20 minutes. While no explicit time trials were run with the old piperidine

34

(the new piperidine came in at the same time we were planning on running the time

trials), time trials with newly purchased piperidine showed no increase in deprotection

after 20 minutes, suggesting that 20 minutes was adequate time for deprotection

(Figure 11).

The other goal of this project was to prove that a solid-phase coupling reaction

using modern coupling reagents and Fmoc amino acids was possible in propylene

carbonate. To this end, a valine-phenylalanine dipeptide was synthesized on H-Rink

amide ChemMatrix resin using PyBOP as the coupling reagent and propylene

carbonate as the primary solvent. Though we were unable to get final data for the

success of this reaction, the substitution assay at each stage of synthesis (after Phe

addition, and after completion) indicates that the coupling was entirely successful

(Figure 12). Since the uncoupled resin does not have a protecting group in it it does not

show up at all in the substitution level assay, but the results of the assay for both the

phenylalanine-resin and the Val-Phe-resin are in the range we would expect from a

successful coupling reaction to the resin.

Overall, even though we were unable to get final characterization of a peptide

synthesized on solid phase in propylene carbonate, our results strongly suggest that it is

a viable option as a greener solvent in peptide synthesis. All of the required reagents

can dissolve in PC, the assay can be completed using PC, and it is possible to make a

dipeptide using PC as the primary solvent.

35

36

Future Directions

Now that it has been demonstrated that a modern solid phase coupling reaction

can take place in propylene carbonate, the next step is to get quantitative data

regarding the success of coupling in PC and compare it to the same reaction in DMF.

The assay has been improved to the point where it is able to give reasonably

reproducible data and could certainly be used for a comparison between the two

solvents. Once the direct comparison has been made, it is vital to adapt the synthesis to

propylene carbonate and attempt to improve yield. Furthermore, it is important to test

propylene carbonate as a solvent for more difficult coupling reactions and ensure that it

is a successful solvent in those as well. The Phe-Val coupling was initially chosen due

to its relative simplicity. By researching and testing more complex coupling reactions in

propylene carbonate, we would aim to prove that PC is as versatile a solvent for peptide

synthesis as DMF. This may also involve the use of coupling reagents besides HBTU

and PyBOP, as other reagents are more suited to specific difficult reactions.

One interesting result from this study was the correlation between an amino

acid’s inability to dissolve in PC, and the presence of an unprotected amine. Further

experimentation and research into why this might be may help to isolate alternative

cosolvents (other than DMF) that could be used to get these amino acids to dissolve.

Finally, it is important to determine whether the final step of peptide synthesis

(orthogonal deprotection and cleavage from the resin with TFA) can be performed in

propylene carbonate. Though this step does use less solvent in comparison to the

repetitive cycle of coupling and deprotection, its successful completion in PC would

37

mean that the entirety of the peptide synthesis process could be performed with PC as

the primary solvent.

38

Experimental Procedures

Substitution Assay for Aminoacylated Polystyrene Resin

A sample of the resin bound peptide was washed with either dichloromethane or

ethanol and dried under vacuum. A 10-20 mg sample of the dried resin was weighed

and mixed with 1.00 or 2.00 mL of 20% piperidine (vol/vol) in DMF. After 20 minutes,

0.200 mL of the solution was removed and diluted to a total volume of 4.00 mL using

either acetonitrile or ethanol. The absorbance at 300 nm was measured using a UV-Vis

Spectrometer (What type). Dibenzofulvene has an extinction coefficient (ε) of 7040 M-

1cm-1 at 300 nm. Thus, by Beer’s Law, for X g of resin, an absorbance of A, and Y mL of

initial deprotection solution, the substitution level of the resin sample is given by the

equation 𝑆𝐿 = (20 ∗ 𝑌 ∗ 𝐴)/(7040 ∗ 𝑋). It is extremely important to use fresh piperidine

in this assay. Initial trials were run in old piperidine, which was discolored yellow, and

caused much more variation in the end results.

Coupling of amino acid to ChemMatrix Resin (general procedure)

H-Rink amide ChemMatrix resin (0.1 to 0.5 g, 0.4 mmol/g) was added to a 10 mL

solid-phase synthesis reaction vessel. Fmoc-protected amino acid (1.1 equivalents) was

dissolved in 3 mL propylene carbonate. PyBOP (1.1 equivalents) was dissolved in 0.5

mL propylene carbonate. Propylene carbonate (2-5 mL) was added to the reaction

vessel to swell the resin. The mixture was shaken for 1 minute and the propylene

carbonate removed by filtration. This procedure was repeated three times. The Fmoc

amino acid solution, PyBOP solution and diisopropylethylamine (2.2 equivalents) were

39

added. The vessel was shaken for one hour. The solution was then removed by

filtration. The propylene carbonate washing procedure (add 2-5 mL of propylene

carbonate, shake for one minute, remove propylene carbonate by filtration) was

repeated four times. If necessary, an aliquot of resin was removed for testing at this

point following the procedure listed below.

Deprotection of Fmoc group from Fmoc-aminoacyl-ChemMatrix resin:

Fmoc-aminoacylated resin (0.5 - 1.0 g) was washed 3-4 times with propylene

carbonate. Then a 20% by volume solution of piperidine in propylene carbonate was

prepared and added to the reaction vessel until it covered the resin. The reaction vessel

was shaken for 20 to 30 minutes to allow the resin to deprotect, and then washed

thoroughly with propylene carbonate to remove the piperidine. This prepared the resin

for either a new coupling or the substitution assay described above.

Solubility Tests

The solubility of 13 Fmoc amino acids in propylene carbonate was determined.

Solubility tests were completed using 0.0001 mols of each amino acid and coupling

reagent. 1.00 mL of propylene carbonate was added to each test tube. Each sample

was given a minute in an ultrasonic bath. Afterwards, the samples that had dissolved

were noted. For the samples that had not dissolved, another 1.00 mL of propylene

carbonate was added, and the samples were again placed in the ultrasonic bath. It was

previously determined that a more dilute solution of the reagents was acceptable.

Again, it was noted which amino acids had dissolved. For the five amino acids that

40

remained undissolved, DMF was added dropwise until either the amino acid dissolved,

or 100 drops were reached. This data was recorded.

NMR

Many of our reagents were old and needed to be tested for purity. In order to do

this, NMR samples containing each reagent of interest dissolved in deuterated

chloroform with TMR were prepared and tested using a 500MHz NMR. The resulting

spectra were scrutinized for possible contaminants.

41

References

(1) Tanabe, S. Epitope Peptides and Immunotherapy. 10.

(2) Zhang, L. Antimicrobial Peptides. Current Biology 6.

(3) Zhang, C.; Xue, X.; Luo, Q.; Li, Y.; Yang, K.; Zhuang, X.; Jiang, Y.; Zhang, J.; Liu, J.; Zou, G.; Liang, X.-J. Self-Assembled Peptide Nanofibers Designed as Biological Enzymes for Catalyzing Ester Hydrolysis. 2014, 8 (11), 9.

(4) Fosgerau, K.; Hoffmann, T. Peptide Therapeutics: Current Status and Future Directions. Drug Discovery Today 2015, 20 (1), 122–128. https://doi.org/10.1016/j.drudis.2014.10.003.

(5) Cromm, P. M.; Spiegel, J.; Grossmann, T. N. Hydrocarbon Stapled Peptides as Modulators of Biological Function. ACS Chem. Biol. 2015, 10 (6), 1362–1375. https://doi.org/10.1021/cb501020r.

(6) Lokhande, A. C.; Gurav, K. V.; Jo, E.; Lokhande, C. D.; Kim, J. H. Chemical Synthesis of Cu 2 SnS 3 (CTS) Nanoparticles: A Status Review. Journal of Alloys and Compounds 2016, 656, 295–310. https://doi.org/10.1016/j.jallcom.2015.09.232.

(7) Fields, G. B. Introduction to Peptide Synthesis. Current Protocols in Protein Science 2001, 26 (1). https://doi.org/10.1002/0471140864.ps1801s26.

(8) Lawrenson, S. B.; Arav, R.; North, M. The Greening of Peptide Synthesis. Green Chem. 2017, 19 (7), 1685–1691. https://doi.org/10.1039/C7GC00247E.

(9) Atherton, E.; Sheppard, R. C. Solid Phase Peptide Synthesis: A Practical Approach; A Practical Approach.

(10) El-Faham, A.; Albericio, F. Peptide Coupling Reagents, More than a Letter Soup. Chem. Rev. 2011, 111 (11), 6557–6602. https://doi.org/10.1021/cr100048w.

(11) Kimmerlin, T.; Seebach, D. ‘100 Years of Peptide Synthesis’: Ligation Methods for Peptide and Protein Synthesis with Applications Toβ-Peptide Assemblies. Journal of

42

Peptide Research 2005, 65 (2), 229–260. https://doi.org/10.1111/j.1399-3011.2005.00214.x.

(12) Palasek, S. A.; Cox, Z. J.; Collins, J. M. Limiting Racemization and Aspartimide Formation in Microwave-Enhanced Fmoc Solid Phase Peptide Synthesis. Journal of Peptide Science 2007, 13 (3), 143–148. https://doi.org/10.1002/psc.804.

(13) Coste, J.; Le-Nguyen, D.; Castro, B. PyBOP: A New Peptide Coupling Reagent Devoid of Toxic By-Product. Tetrahedron Letters 1990, 31 (2), 205–208.

(14) Dunn, B.; Pennington, M.; Byrnes, M. Proceedures to Improve Difficult Couplings. In Peptide Synthesis Protocols; 1994; pp 1–16.

(15) Albericio, F. Orthogonal Protecting Groups for Nα-Amino and C-Terminal Carboxyl Functions in Solid-Phase Peptide Synthesis. Peptide Science 2000, 55 (2), 123–139. https://doi.org/10.1002/1097-0282(2000)55:2<123::AID-BIP30>3.0.CO;2-F.

(16) Al Musaimi, O.; Jad, Y. E.; Kumar, A.; El-Faham, A.; Collins, J. M.; Basso, A.; de la Torre, B. G.; Albericio, F. Greening the Solid-Phase Peptide Synthesis Process. 2-MeTHF for the Incorporation of the First Amino Acid and Precipitation of Peptides after Global Deprotection. Org. Process Res. Dev. 2018, 22 (12), 1809–1816. https://doi.org/10.1021/acs.oprd.8b00335.

(17) Deepshikha Verma; Pillai V N R; Giriraj Tailor. Role of Capping in Peptide Synthesis. ijrps 2020, 11 (4), 5225–5228. https://doi.org/10.26452/ijrps.v11i4.3134.

(18) Albericio, F. Encyclopedia of Reagents for Organic Synthesis; 2001.

(19) Anastas; Beach. History of Green Chemistry | Center for Green Chemistry & Green Engineering at Yale https://greenchemistry.yale.edu/about/history-green-chemistry (accessed Jan 23, 2021).

(20) Behrendt, R.; Huber, S.; Martí, R.; White, P. New t -Butyl Based Aspartate Protecting Groups Preventing Aspartimide Formation in Fmoc SPPS: NEW t -BUTYL BASED ASPARTATE PROTECTING GROUPS FOR FMOC SPPS. J. Pept. Sci. 2015, 21 (8), 680–687. https://doi.org/10.1002/psc.2790.

(21)

43

Chaudhari, M. I.; Muralidharan, A.; Pratt, L. R.; Rempe, S. B. Assessment of Simple Models for Molecular Simulation of Ethylene Carbonate and Propylene Carbonate as Solvents for Electrolyte Solutions. In Modeling Electrochemical Energy Storage at the Atomic Scale; Korth, M., Ed.; Topics in Current Chemistry Collections; Springer International Publishing: Cham, 2018; pp 53–77. https://doi.org/10.1007/978-3-030-00593-1_3.

(22) Cherkupally, P.; Ramesh, S.; de la Torre, B. G.; Govender, T.; Kruger, H. G.; Albericio, F. Immobilized Coupling Reagents: Synthesis of Amides/Peptides. ACS Comb. Sci. 2014, 16 (11), 579–601. https://doi.org/10.1021/co500126y.

(23) Declerck, V.; Nun, P.; Martinez, J.; Lamaty, F. Solvent-Free Synthesis of Peptides. Angewandte Chemie International Edition 2009, 48 (49), 9318–9321. https://doi.org/10.1002/anie.200903510.

(24) The Origins of Organic Chemistry https://chem.libretexts.org/Courses/Sacramento_City_College/SCC%3A_Chem_420_-_Organic_Chemistry_I/Text/01%3A_Introduction_and_Review/1.01%3A_The_Origins_of_Organic_Chemistry (accessed Jan 23, 2021).

(25) Frérot, E.; Coste, J.; Pantaloni, A.; Dufour, M.-N.; Jouin, P. PyBOP® and PyBroP: Two Reagents for the Difficult Coupling of the α,α-Dialkyl Amino Acid, Aib. Tetrahedron 1991, 47 (2), 259–270. https://doi.org/10.1016/S0040-4020(01)80922-4.

(26) Gescher, A. Metabolism of N,N-Dimethylformamide: Key to the Understanding of Its Toxicity. Chem. Res. Toxicol. 1993, 6 (3), 245–251. https://doi.org/10.1021/tx00033a001.

(27) Han, S.-Y.; Kim, Y.-A. Recent Development of Peptide Coupling Reagents in Organic Synthesis. Tetrahedron 2004, 60 (11), 20.

(28) Recent Development of Coupling Reagents.Pdf.

(29) Hu, L.; Xu, S.; Zhao, Z.; Yang, Y.; Peng, Z.; Yang, M.; Wang, C.; Zhao, J. Ynamides as Racemization-Free Coupling Reagents for Amide and Peptide Synthesis. J. Am. Chem. Soc. 2016, 138 (40), 13135–13138. https://doi.org/10.1021/jacs.6b07230.

(30)

44

Ingenito, R.; Bianchi, E.; Fattori, D.; Pessi, A. Solid Phase Synthesis of Peptide C-Terminal Thioesters by Fmoc/ t -Bu Chemistry. J. Am. Chem. Soc. 1999, 121 (49), 11369–11374. https://doi.org/10.1021/ja992668n.

(31) Kvsrg, P. APPLICATIONS OF PEPTIDE COUPLING REAGENTS – AN UPDATE. 8 (1), 12.

(32) Li, C.-J.; Trost, B. M. Green Chemistry for Chemical Synthesis. 6.

(33) Metcalfe, I. S.; North, M.; Pasquale, R.; Thursfield, A. An Integrated Approach to Energy and Chemicals Production. Energy Environ. Sci. 2010, 3 (2), 212–215. https://doi.org/10.1039/B918417A.

(34) Nehlig, E.; Motte, L.; Guénin, E. Nano-Organocatalysts Synthesis: Boc vs Fmoc Protection. Catalysis Today 2013, 208, 90–96. https://doi.org/10.1016/j.cattod.2012.10.027.

(35) Protein Biology Resource Library. Peptide Synthesis - US //www.thermofisher.com/us/en/home/life-science/protein-biology/protein-biology-learning-center/protein-biology-resource-library/pierce-protein-methods/peptide-synthesis.html (accessed Feb 11, 2021).

(36) Riester, D.; Wiesmüller, K.-H.; Stoll, D.; Kuhn, R. Racemization of Amino Acids in Solid-Phase Peptide Synthesis Investigated by Capillary Electrophoresis. Anal. Chem. 1996, 68 (14), 2361–2365. https://doi.org/10.1021/ac9511511.

(37) Sheehan, J. C.; Hlavka, J. J. The Use of Water-Soluble and Basic Carbodiimides in Peptide Synthesis. J. Org. Chem. 1956, 21 (4), 439–441. https://doi.org/10.1021/jo01110a017.

(38) Sigma. Peptide Coupling Reagents Selection Guide https://www.sigmaaldrich.com/technical-documents/articles/chemistry/peptide-coupling-reagents-selection-guide.html (accessed Jan 29, 2021).

(39) Stawikowski, M.; Fields, G. B. Introduction to Peptide Synthesis. Curr Protoc Protein Sci 2002, CHAPTER, Unit-18.1. https://doi.org/10.1002/0471140864.ps1801s26.

(40) About Green Chemistry. Beyond Benign.

45

(41) Resins for Solid Phase Peptide Synthesis - Core Resins. AAPPTEC.

(42)

Fontenot JD, Ball JM, Miller MA, David CM, Montelaro RC. A survey of potential problems and quality control in peptide synthesis by the fluorenylmethoxycarbonyl procedure. Pept Res. 1991 Jan-Feb;4(1) 19-25. PMID: 1802234.

(43) Prat, Denis, John Hayler, and Andy Wells. "A survey of solvent selection guides." Green Chemistry 16, no. 10 (2014): 4546-4551.


Recommended