calendar

Research in Biology

Utilizing Environmental DNA for the Detection of Invasive Goldfish (Carassius Auratus) and Koi (Cyprinus rubrofuscus) in Southeastern Massachusetts Water System 

By: Mia Oliviera

I am happy to report that this semester’s research has been a huge, fun, learning opportunity, and I am very appreciative of the OUR’s grant to assist me in my work this Spring. In March, I spent my spring break taking water samples and negative controls at all ten sites across Southeastern Massachusetts. These samples were then brought to the UMD campus to be filtered over the following few weeks within Dr. Drew’s genetics lab. Samples were filtered using a vacuum pump setup under a sterile hood, and bleach was used to ensure no cross-contamination between samples.

Starting in early April. I utilized the DNeasy Blood and Tissue Kit and the Powersoil PowerLyzer kit to optimize an eDNA extraction method that best suits my extraction needs. My two positive control samples (goldfish and koi ponds) were extracted with both kits to determine which would produce the best product. It was found that the Powersoil PowerLyzer kit allowed the filter and cells to break down best and produce a clearly visible PCR product. I determined the best extraction method was to utilize the DNEasy Powersoil PowerLyzer kit for the remainder of my samples, with a few modifications to the listed protocol.

Once all samples and negative controls were extracted, I ran a PCR utilizing Dr. Drew’s genetics lab resources and using universal fish DNA primers that amplify all fish species. I chose to use fish primers instead of my species-specific primers to ensure the filtering, extraction, and PCR methods captured fish eDNA. The PCR products were run in an electrophoresis gel and imaged. The first PCR I ran produced a lot of DNA bands but showed signs of potential contamination within the PCR step. I chose to do a second PCR while under a sterile hood, which produced a cleaner product and a few samples containing DNA, but did show some amplification of fish DNA within my negative control samples. Dr. Drew and I worked through troubleshooting methods and discussed ways we could optimize protocols for the Fall semester and ways to use different primers, as well as my species-specific primers. I decided to do more research on eDNA sampling, extraction, and PCR methods over the summer to find the best protocols for my research in the Fall.

I have learned so much about the world of eDNA over the last few months. I find the troubleshooting aspect not to be a roadblock, but a huge opportunity to problem-solve and optimize my own protocols. eDNA research is still in its early stages, but it has so much potential in its application. I am very excited to continue my research on this in the Fall, and once again thank the OUR for this funding opportunity.

(Sampling at Snows Pond, Rochester, MA)

Research in Chemistry and Biochemistry

Chemical Profiling and Evaluation of Antioxidant and Neurodegenerative Enzyme Inhibitory Potential of Flavonol and Flavonol Glycoside Fractions of Cranberry Pomace and Fruit Extracts

By: Ryley Thatcher
Abstract

The large cranberry (Vaccinium macrocarpon) is widely known for its antioxidant potential due to its potent polyphenolic chemical constituents, such as anthocyanins, flavonols, phenolic acids, and proanthocyanidins. The composition of polyphenolic compounds within a cranberry sample varies with different factors such as growth stage, growth region, and the cultivar of the berry, which successively affect antioxidant capacity. This study aimed to analyze and compare the phytochemical composition and antioxidant activity of five new 2024 cranberry cultivars (Ruby Star, Mullica Queen, Pilgrim King, Welkers, and Haines) grown by the Cranberry Station in Wareham, MA. A small-scale extraction method was optimized, and antioxidant activity was determined by colorimetric assays, including the DPPH and FRAP assays. Results from all cultivars showed a dose-dependent response for both assays. Welkers had the best DPPH scavenging activity (IC50 =104 μg/mL ∓ 17), and Pilgrim King had the best reducing power based on FRAP (IC50 =184 μg/mL ∓ 11). The total phenolic content and proanthocyanidin (PAC) content of samples were measured by the Folin-Ciocalteu and DMAC assays, respectively. Haines had the highest total phenolic content, and Mullica Queen had the highest PAC content. HPLC-DAD analysis will be conducted to identify and quantify various polyphenolic compounds in the samples. Results will be shared with the Cranberry Station and could potentially inform future agricultural practices and dietary recommendations.

Introduction

Cranberries consist of various bioactive compounds known for their health benefits and have been used in traditional medicine for their antioxidant, anti-inflammatory, and antimicrobial properties. The large cranberry (Vaccinium macrocarpon), also known as the American Cranberry, is native to North America. In the United States, the large cranberry is most commonly grown in New Jersey, Massachusetts, Oregon, and Wisconsin. Beyond North America, V. macrocarpon is also extensively cultivated globally in locations such as Canada, Europe, and South America [1]. While other species of cranberry also contain significant amounts of nutritional phytochemicals, in the modern marketplace, the large cranberry is considered an important success story of the functional foods industry due to its notably high concentration of antioxidant compounds [2].

A specific constituent of cranberries largely associated with their high antioxidant activity is a class of compounds known as polyphenols. The polyphenolic contributors in cranberries are primarily flavonoids, including flavonols, proanthocyanidins, anthocyanins, phenolic acids, and benzoates [3]. The phenolic compounds in cranberries achieve antioxidant effects by scavenging the free radicals of harmful ROS (reactive oxygen species) that, when accumulated in the human body, can be linked to illnesses such as cancers and neurodegenerative diseases [1].

Figure 1. Cranberry Polyphenols: anthocyanins, phenolic acids, flavonols, and phenolic acids

The phytochemical composition of these polyphenols is influenced by multiple factors: cultivar, genotype, growing season, ripening, processing, and storage of cranberry fruit [4]. A cultivar’s influence on the phytochemical composition in cranberries is particularly substantial, given that each cultivar is a specific variety of cranberry, bred for its desirable traits. It is crucial to understand the impact specific cultivars have on the content of these phenolic compounds to maximize the gain of nutritional benefits from cranberries and for any potential medicinal applications in the future. This research will aim to compare the phytochemical composition and antioxidant activity of five Massachusetts Cranberry Cultivars collected on October 2nd, 2024, from the Cranberry Station in Wareham, MA. Results from this project will be shared with the cranberry station to influence future growing strategies and dietary recommendations.

Figure 2. Cultivar Breeding Map (not pictured: Ruby Star: Hy red x Bergman hybridPilgrim King: Pilgrim x GH2 hybrid)

MATERIALS AND METHODS

Sample Collection

Several cultivars of cranberry fruit were collected from the cranberry bogs at the Cranberry Station in Wareham, MA, in October 2024. After collection, samples were flash-frozen with liquid nitrogen and stored at -20°C until further use.

Sample Preparation and Extraction

Flash-frozen cranberry fruit samples from each cultivar were freeze-dried and ground into a fine powder. One gram of cranberry powder per cultivar was dissolved in 10 mL of a 40:40:19.9:0.1 acetone:methanol: formic acid: DI water v/v solvent. The resultant solution was mixed by a stir plate for one hour, followed by ultrasonication for 30 minutes. After mixing, the solutions were stored in the fridge for 24 hours. The next day, suction filtration was performed twice; the first time suction filtration was performed, the residue was re-dissolved in 5 mL of extraction solvent. For each filtration, the filtrate was kept as the liquid extract. Extracts were stored in the freezer until further antioxidant analysis, and the extraction process was performed in triplicate for each cultivar.

Folin-Ciocalteu Assay

The total phenolic content of each cultivar was measured by the Folin-Ciocalteu Assay according to Ainsworth et. al [5] with modification. 20 mg of cranberry powder from each cultivar was dissolved in 2 mL of ice-cold 95% v/v methanol, resulting in a 10 mg/mL sample concentration. Samples were mixed by vortexing and ultrasonication for 30 minutes, and then left in the fridge overnight. The next day, 10% v/v Folin-Ciocalteu reagent was prepared and kept in the dark until plating. Samples from the previous day were centrifuged at 13,000 g for 5 minutes, and the supernatant was used for analysis. 40 μl of each sample was transferred to an Eppendorf tube, and 95% v/v methanol was used as the blank. 800 μl of Folin reagent was transferred to each tube, and samples were left to stand for 5 minutes. 800 μl of 700 mM Na2CO3 was added to each tube and mixed by vortexing. Lastly, 360 μl of distilled water was added to each tube, reaching a final volume of 2 mL. Samples were incubated in the dark for 2 hours at room temperature. After incubation, 200 μl of each sample was plated in triplicate on a clear 96-well microplate. Once samples were plated, absorbance was read at 760 nm using a UV-Spectramax plate reader. Total phenolic content was determined as gallic acid equivalents by comparison to a gallic acid standard curve.

DMAC Assay

The total proanthocyanidin (PAC) content of each cultivar was determined by the DMAC (P-Dimethylaminocinnamaldehyde) assay as previously described by Feliciano et. al [8] with modifications. 50 milligrams of samples from each cultivar were dissolved in 1 mL of a PAC (75%/24.5%/0.5% acetone/water/acetic acid) extraction solution in triplicate. To mix the samples, they were sonicated for five minutes and placed on a Line-lab Orbital Shaker for one hour at 250 rpm. Once the samples were mixed, they were kept in the freezer overnight. The next day, samples were centrifuged at 3,000 rpm for five minutes, and the supernatant was kept for plating. DMAC reagent was prepared by dissolving 0.05 grams of DMAC in 50 mL of a 75%, 12.5%, 12.5% v/v/v ethanol/hydrochloric acid/DI water solution. 5 μl aliquots of each cultivar sample were plated in triplicate on a clear 96-well microplate, followed by the addition of 65 μl PAC extraction solution to each well. 70 μl of PAC extraction solution was plated in triplicate as the blank.= 100 μl of DMAC reagent was added to each well to reach a final volume of 170 μl, and the plate was incubated for 30 minutes. Absorbance was read at 640 nm on the UV Spectramax plate reader. Absorbances were compared to a PAC standard curve to express results as mg PAC per g DW.

DPPH Free Radical Scavenging Antioxidant Activity Assay

The free radical scavenging antioxidant activity of each cranberry sample was evaluated using the 2,2-diphenyl-1-picrylhydrazyl (DPPH) free radical scavenging assay as previously described by Liu et al [6]. A 100 μg/mL DPPH solution was prepared using methanol. 75 μl of cranberry extracts were diluted with 4.925 mL of methanol to achieve a target stock concentration of 1000 μg/mL. A two-factor serial dilution was performed for each cultivar extract from 1000 μg/mL to 31.25 μg/mL. For this assay, different 96-well plates were used for the different cultivar samples, resulting in the use of five different microplates. 100 μl of each concentration was plated in triplicate, followed by the addition of 25 μl of DPPH reagent. 125 μl of methanol was plated in triplicate as the blank. Once all samples were plated, they were incubated for 30 minutes at room temperature in the dark, and absorbance was read at 517 nm by the Spectramax plate reader. Percent inhibition was calculated using the following equation:

% Inhibition= A0-A1/A0 *100%

Where A0 was the absorbance of the control solution with no radical, and A1 was the absorbance
of the sample at 517 nm. Percent Inhibition results were converted to IC50 values by OriginLab Pro software.

FRAP Ferric Reducing Antioxidant Power Assay

A method to examine the ferric reducing antioxidant power of each cranberry sample was developed by previous FRAP procedures with modifications [9][10]. A 10:1:1 acetate buffer, tripyridyltriazine (TPTZ), FeCl3 FRAP working solution was prepared as follows: 50 mL of 300 mM pH 3.6 acetate buffer, 5 mL of 10 mM TPTZ solution in 40 mM HCl, and 5 mL of 20 mM FeCl3 were added to a 250 mL Erlenmeyer flask. The FRAP working solution must be used within three hours of preparation for accurate measurement. Sample preparation was the same as performed for the DPPH assay, obtaining a stock target concentration of 1000 μg/mL to perform a two-factor serial dilution down to 31.25 μg/mL. Similar to the DPPH assay procedure, five separate 96-well microplates were used for each cultivar. 20 μl of each concentration was pipetted into the microplate in triplicate, and methanol was used for the blank. 180 μl of FRAP solution was transferred to each well and mixed by pipette. 200 μl of methanol was plated in triplicate as the blank. Samples were incubated in the dark at room temperature for 30 minutes, and the absorbance was read at 593 nm by the Spectramax microplate reader. The percent reducing power of each sample was calculated by the following equation:

FRAP% = Sample Absorbance – Blank Absorbance/ Sample Absorbance x 100

Percent reducing power results were converted to IC50 values by the Origin Lab Pro software.

RESULTS + DISCUSSION 

Total Phenolic Content 

The objective of this project was to analyze and compare the antioxidant activity of five new 2024 cranberry cultivars, Mullia Queen, Ruby Star, Welkers, Haines, and Pilgrim King, grown by the Cranberry Station in Wareham, MA. The first parameter examined was the total phenolic content by the Folin-Ciocalteu assay. The Folin assay measures the ability of polyphenols present in the cranberry fruit samples to reduce the Folin Reagent, which is an oxidant that consists of a combination of phosphotungstic and phosphomolybdic acids. A color change from yellow to blue occurs at 760 nm, signifying the completion of the reduction reaction. Results showed that cultivar Haines had the highest total phenolic content (32.5土2.5 mg gallic acid/g dry powder), while Mullica Queen (MQ) had the lowest total phenolic content of the five cultivars (25.7 土4.4 mg gallic acid/g dry powder). Cultivars Welkers (30.4 土2.2 mg gallic acid/ g dry powder), Ruby Star (29.1 土2.1 mg gallic acid/ g dry powder), and Pilgrim King (27.8 土 0.51 mg gallic acid/ g dry powder) fell in the middle. Error bars indicate overlapping standard deviations between some cultivars, indicating that the total phenolic content between these cultivars was not completely significantly different.

Figure 3. Comparison of Total Phenolic Content by Folin Assay

Total Proanthocyanidin Content

Proanthocyanidins are widely regarded in the industry for their urinary tract health benefits. PACS are specifically known for their potential in the prevention and treatment of UTIs due to their anti-adhesion capabilities that can prevent E. coli bacteria from sticking to the walls of the bladder and urinary tract. The DMAC assay was used to quantify total soluble PAC content in each cranberry cultivar sample. The DMAC reagent reacts with PACS through a condensation mechanism between the aldehyde of the DMAC and the C8 of the PAC’s terminal unit, which is read at an absorbance of 640 nm. The DMAC assay is a more accurate colorimetric detection of PACs due to the specificity of the reaction that minimizes interference from other polyphenolic compounds that may be present in the samples.

Figure 4. Comparison of Total PAC Content

Mullica Queen was found to have the highest PAC content (78.7 土 2.7 mg/ g dry powder) of the five cultivars analyzed, and Haines was found to have the lowest PAC content (39.5 土 2.7 mg/ g dry powder). Pilgrim King, Ruby Star, and Welkers were in the middle of the range with a PAC content of 64.7 土 4.3 mg/ g dry powder, 62.2 土 1.7 mg/g dry powder, and 43.6 土 2.6 mg/g dry powder, respectively. Error bars are relatively small, indicating consistent results across replicates. Interestingly, even though Haines was the highest in total phenolic content, the trend shows Haines to have the lowest PAC content, which could suggest the Haines cultivar is higher in other polyphenolic constituents like anthocyanins, phenolic acids, and flavonols. MQ was the lowest in total phenolic content but highest in PAC content, indicating that it has a high PAC: phenolic compound ratio. Since one of the goals of cultivating Mullica Queen was to produce a cultivar with high PAC content, MQ being high in PAC content was expected.

Free Radical Scavenging Antioxidant Activity

The DPPH assay was used to determine the free radical scavenging ability of the cranberry fruit samples. The assay measures the antioxidant activity by measuring the cranberry polyphenols’ ability to scavenge the DPPH radical reagent. The neutralization of the radical is signified by a visible color change from purple to yellow at 517 nm.

Cultivar IC50 (μg/mL)
Welkers 105 土 17
Ruby Star 144 土 24
Pilgrim King 172 土 42
Haines 227 土 32
Mullica Queen 330 土 57

Table 1. Comparison of Free Radical Scavenging by IC50 Value

Free radical scavenging results were reported as IC50 values, which measure the concentration of the sample that would be required to inhibit 50% of the DPPH radicals. IC50 is inversely related to antioxidant activity, implying that a lower IC50 specifies a stronger antioxidant ability. Based on the data presented in Table 1 and Figure 5, Welkers had the highest free radical scavenging capacity (IC50= 105 μg/mL 土 17) followed by Ruby Star (144 μg/mL土 24), Pilgrim King (172 μg/mL 土 42), and Haines (330 μg/mL 土 57). Mullica Queen had the weakest free radical scavenging ability (IC50=330 μg/mL 土 57).

Figure 5. Free Radical Scavenging Capacity Measured by DPPH Assay

The results show that even though Mullica Queen had the highest total PAC content, it had the lowest free radical scavenging capacity. Welkers had the best free radical scavenging activity, and it performed moderately in both total phenolic and total PAC assays. This could suggest that some types of polyphenols contribute more to free radical scavenging capacity than others. For example, the MQ results propose that PACS may not have a significant effect on the antioxidant activity for this reaction. Figure 6 shows an example of how, for all samples, antioxidant activity decreased with concentration.

Figure 6. Example of Dose Response DPPH Results

Ferric Reducing Antioxidant Power

Another mechanism of antioxidant activity was measured by the FRAP assay, which measured the ferric reducing antioxidant power of each cultivar. The FRAP assay quantifies the ability of antioxidants in a sample to reduce Fe3+ to Fe2+ in an acidic environment. The reaction produces a blue-colored complex, which has a wavelength of 593 nm. The intensity of the consequent blue color correlates with the reducing power of the sample.

Cultivar IC50 (μg/mL)
Pilgrim King 184 土 11
Mullica Queen 234 土 6
Welkers 273 土 25
Ruby Star 303 土 21
Haines 377 土 37

Table 2. Comparison of Ferric Reducing Power Measured by FRAP Assay

Similarly to the DPPH results, ferric reducing power was expressed as IC50 values. As seen in Table 2 and Figure 7, Pilgrim King had the lowest IC50 value (184 μg/mL 土 11), indicating it had the highest ferric reducing power out of the five cultivars. Mullica Queen and Welkers had relatively similar IC50 values (234 μg/mL 土 6 and 273 μg/mL 土 25, respectively), followed by Ruby Star (303 μg/mL 土 21). Haines had the lowest ferric reducing power of the samples with an IC50 of 377 土 37 μg/mL).

Figure 7. Ferric Reducing Power Measured by FRAP Assay

Interestingly, some of the FRAP trends disagree with the DPPH results. For example, Mullica Queen had the weakest free radical scavenging activity in the DPPH assay, but had the second-best ferric reducing power as seen in Figure 7. These discrepancies between the DPPH and FRAP results highlight that the antioxidant activity performance of samples depends on the reaction mechanism being measured, and different cultivars may be better or worse depending on the parameter being examined. Figure 8 shows an example of the dose-dependent response data that was obtained for each cultivar, demonstrating how, as concentration decreases, reducing power decreases.

Figure 8. Example of Dose-Dependent FRAP Data

CONCLUSIONS

This project highlights that the cranberry cultivar has a substantial impact on polyphenolic composition and antioxidant activity between cranberry fruit samples. A variety of colorimetric assays were utilized to measure parameters of each cultivar sample, such as total phenolic content, total proanthocyanidin content, free radical scavenging capability, and ferric reducing power. Cultivar Haines was found to have the highest total phenolic content, while Mullica Queen was found to have the lowest phenolic content of the five cultivars. Oppositely, MQ was found to have the highest PAC content, and Haines was found to have the lowest PAC content. DPPH IC50 data show Welkers to have the overall strongest radical scavenging antioxidant capacity, and Mullica Queen to have the least free radical scavenging capacity. FRAP IC50 values demonstrated Pilgrim King to have the highest percent ferric reducing power, and cultivar Haines had the weakest ferric reducing power. Inconsistencies in trends between assays indicate that antioxidant activity differs based on the assay or mechanism that is being measured, so different cultivars may be desired for different purposes. As the present work provides an initial idea on the different cultivars’ polyphenolic composition and antioxidant activity, future work for this project would be to quantify and compare the polyphenolic composition of anthocyanins, phenolic acids, flavonols, and PACS present in the samples by HPLC-DAD analysis.

ACKNOWLEDGMENTS

The authors would like to acknowledge the support of the UMass Cranberry Health Research Center, Giverson Mupambi at the UMass Cranberry Station, and the UMD Office of Undergrad Research for this research project.

REFERENCES

(1)Nemzer, B. V.; Al-Taher, F.; Yashin, A.; Revelsky, I.; Yashin, Y. Cranberry: Chemical
Composition, Antioxidant Activity and Impact on Human Health: Overview. Molecules 2022, 27(5), 1503. https://doi.org/10.3390/molecules27051503.

(2)Brown, P.; Turi, C.; Shipley, P.; Murch, S. Comparisons of Large (Vaccinium macrocarpon Ait.) and Small (Vaccinium oxycoccos L., Vaccinium Vitis-Idaea L.) Cranberry in British Columbia by Phytochemical Determination, Antioxidant Potential, and Metabolomic Profiling with Chemometric Analysis. Planta Medica 2012, 78 (06), 630 640.https://doi.org/10.1055/s-0031-1298239.

(3)Balawejder, M.; Piechowiak, T.; Kapusta, I.; Chęciek, A.; Matłok, N. In Vitro Analysis of
Selected Antioxidant and Biological Properties of the Extract from Large-Fruited Cranberry Fruits. Molecules 2023, 28 (23), 7895. https://doi.org/10.3390/molecules28237895.

(4)Xue, L.; Otieno, M.; Colson, K.; Neto, C. Influence of the Growing Region on the
Phytochemical Composition and Antioxidant Properties of North American Cranberry Fruit
(Vaccinium Macrocarpon Aiton). Plants 2023, 12 (20), 3595–3595.https://doi.org/10.3390/plants12203595.

(5)Ainsworth, E. A.; Gillespie, K. M. Estimation of Total Phenolic Content and Other Oxidation Substrates in Plant Tissues Using Folin–Ciocalteu Reagent. Nature Protocols 2007, 2 (4),875–877. https://doi.org/10.1038/nprot.2007.102.

(6)Liu, J.; Zhao, J.; Dai, Z.; Lin, R.; Wang, G.; Ma, S. A Pair of New Antioxidant Phenolic Acid Stereoisomers Isolated from Danshen Injection (Lyophilized Powder). Molecules 2014, 19 (2),1786–1794. https://doi.org/10.3390/molecules19021786.

(7)Çelik, H.; Özgen, M.; Serçe, S.; Kaya, C. Phytochemical Accumulation and Antioxidant
Capacity at Four Maturity Stages of Cranberry Fruit. Scientia Horticulturae 2008, 117 (4),345–348. https://doi.org/10.1016/j.scienta.2008.05.005.

(8) Feliciano, R. P.; Shea, M. P.; Shanmuganayagam, D.; Krueger, C. G.; Howell, A. B.; Reed, J. D. Comparison of Isolated Cranberry (Vaccinium Macrocarpon Ait.) Proanthocyanidins to Catechin and Procyanidins A2 and B2 for Use as Standards in the
4-(Dimethylamino)Cinnamaldehyde Assay. Journal of Agricultural and Food Chemistry 2012,60 (18), 4578–4585. https://doi.org/10.1021/jf3007213.

(9)Klavins, L.; Perkons, I.; Mezulis, M.; Viksna, A.; Klavins, M. Procyanidins from Cranberry Press Residues—Extraction Optimization, Purification and Characterization. Plants 2022, 11(24), 3517. https://doi.org/10.3390/plants11243517.

(10) Tomasina, F.; Carabio, C.; Celano, L.; Thomson, L. Analysis of Two Methods to Evaluate Antioxidants. Biochemistry and Molecular Biology Education 2012, 40 (4), 266–270.
https://doi.org/10.1002/bmb.20617.

Research in Biology and Biochemistry

Chemical Profiling and Evaluation of Antioxidant and Neurodegenerative Enzyme Inhibitory Potential of Flavonol and Flavonol Glycoside Fractions of Cranberry Pomace and Fruit Extracts

By: Elena De Pra
Abstract

The imbalance between reactive oxygen species and antioxidant defense is often implicated in neuronal damage associated with neurodegenerative diseases like Alzheimer’s and Parkinson’s. Quercetin, a naturally occurring flavonoid, has been extensively studied for its promising neuroprotective potential due to its ability to effectively combat oxidative stress as a potent antioxidant. Typically found in the standard diet as quercetin glycosides in foods such as cranberries, quercetin is known for its free radical scavenging abilities, though its poor bioavailability limits its therapeutic effectiveness. This research aimed to investigate the antioxidant and enzymatic inhibitory potential of two fractions isolated from cranberry pomace and cranberry fruit extract containing quercetin and quercetin-3-galactoside, respectively. The extracts were profiled for polyphenolic compounds, and the desired flavonoids were isolated and confirmed using high-performance liquid chromatography (HPLC). Antioxidant activity was assessed using DPPH, ABTS free radical scavenging, and FRAP reducing power assays. Results demonstrate that the aglycone quercetin fraction exhibited higher antioxidant capacity than the quercetin-3-galactoside fraction. The quercetin fraction also demonstrated strong inhibitory activity against acetylcholinesterase. Future research will evaluate the inhibitory effect of these fractions to inhibit other enzymes associated with neurodegenerative diseases, including butyrylcholinesterase and monoamine oxidase A/B. Results from this study aim to shed light on how glycosylation may influence therapeutic potential with respect to inhibitory effects, to link antioxidant activity to neuroprotection.

Introduction 

Neurodegenerative disorders represent a critical public health crisis affecting millions of individuals regardless of age, sex, education, or income. The World Health Organization (WHO) estimates 6.8 million people die per year of neurological disorders1. The significance of this issue grows as human life spans extend, making elderly populations increasingly vulnerable to neurological diseases like Alzheimer’s and Parkinson’s. This topic is sensitive not only to directly affected individuals, but also to their families, caregivers, and society at large, who fear the uncertainties of aging. Modern therapeutics focus on symptom management, as no treatments currently exist that reverse neuronal death or effectively delay progression.

Natural products have been used for thousands of years in traditional medicine for their biological and pharmacological properties. Natural product-derived compounds show promising potential in drug development for their role in treating neurodegenerative disorders, particularly attributed to their polyphenolic content2. Polyphenolic compounds are proven to have neuroprotective effects, notably through their capacity to neutralize reactive oxygen species, characteristic of their antioxidant activity2. Consuming these naturally occurring compounds, rich in various fruits and vegetables, is a preferred alternative over pharmaceutical drugs as it has both restorative and preventative potential without adverse effects (Fig. 1).

Fig. 1. Schematic Illustration of Neuroprotective Flavonoids Against Oxidative Stress-induced Neurodegeneration

Flavonoids, a class of polyphenolic compounds, exhibit significant cognitive potential, partly through their ability to enhance antioxidant defenses and protect against oxidative stress3. Quercetin, a naturally occurring flavonoid, has been extensively researched for its neuroprotective and anti-inflammatory properties and promising tool in treating neurodegenerative diseases such as Alzheimer’s and Parkinson’s4. Quercetin-3-galactoside is the predominant glycosidic form of quercetin present in cranberries5. The main structural difference lies in the sugar moiety attached in place of the 3’ OH group connected by a glycosidic bond on the central-C-ring of the core quercetin structure (Fig. 2).

Fig. 2. Quercetin and Quercetin-3-galactoside Chemical Structure

Quercetin glycosides are more commonly found in derivatives than free form quercetin6. Glycosides are more water-soluble than their respective non-sugared aglycones and have generally been reported to possess greater bioavailability in vivo7. Although glycosides often have reduced antioxidant activity, their structural differences can enhance compound stability and absorption into the bloodstream, highlighting their potential as a preferred alternative in targeting specific drug properties7.

Although quercetin has been investigated for its inhibitory effects on neurodegenerative enzymes, there is limited research investigating quercetin-3-galactoside’s potential as an inhibitor of neurodegenerative-related enzymes. Acetylcholinesterase (AChE) is an enzyme responsible for hydrolyzing acetylcholine, a critical neurotransmitter that plays a role in memory, learning, and attention8. Butyrylcholinesterase (BChE) is a nonspecific enzyme that catalyzes the hydrolysis of choline and non-choline esters, including acetylcholine9. Low levels of acetylcholine are characteristic of individuals affected by Alzheimer’s disease8. Monoamine oxidase (MAO) exists as two isoforms: MAO-A, which oxidizes serotonin, norepinephrine, and epinephrine, and MAO-B responsible for oxidizing dopamine10. Monoamine oxidases are relevant to Parkinson’s disease, a neurological disease impacted by reduced levels of dopamine, norepinephrine, and serotonin11. Targeting these enzymes and discovering these compounds’ inhibitory potential could highlight their therapeutic relevance.

Comparing isolated fractions of quercetin and its glycosides’ antioxidant capacity and neuroprotective potential is relevant to the development of modern therapeutics that can be used in combination with other potential natural compounds to address root mechanisms of disease. It is important to explore the relationship between flavonoids and their glycosides and how structural differences can impact antioxidant and neuroprotective properties to optimize treatment. In addition to its potential in modern therapeutics, findings from this study, if successful, could influence dietary recommendations to reduce the risk of onset of neurodegenerative diseases12.

Experimental 

Preparation of Crude Extracts 

The Mullica Queen (MQ) cranberry cultivar was collected from cranberry bogs at the Cranberry Station in Wareham, MA, in September 2023. Cranberry pomace samples, consisting of a mixed variety of cultivars, were obtained from Ocean Spray. All samples were flash-frozen in liquid nitrogen and stored at -20°C. Samples were ground, lyophilized, and further ground into a fine powder. Dried fruit powder (40 g) was dissolved (400 mL) in an extraction solvent composed of acetone/methanol/distilled water/formic acid (40:40:19:1 v/v). The mixture was stirred for 1 hour, sonicated for 30 minutes, and refrigerated overnight. The following day, the sample was vacuum filtrated, and the solid residue was re-extracted using half the original solvent volume (200 mL). The same process was repeated the following day with half the original solvent amount for a total of three extractions. All filtrates were combined, rotary evaporated, frozen, and
lyophilized for further analysis.

Isolation of Concentrated Fruit Extracts 

A Diaion HP-20 chromatography column was prepared for the purification of polyphenolic compounds from the Mullica Queen (MQ) extract. A glass chromatography column was packed with clean sand and glass wool. The resin was pretreated by activating with methanol for 15 minutes and washing with distilled water prior to being transferred into the column, allowing the resin to settle. Crude extract (20 g) was dissolved in a minimal volume of methanol/formic acid (99.9:0.1v/v) and was loaded onto the column and allowed to absorb into the packing. After 15 minutes had elapsed, the elution of free sugars was initiated using distilled water until the eluate appeared colorless; colorless eluates were discarded. Colored bands containing polyphenols and terpenoids were eluted using methanol/formic acid (99.9: 0.1 v/v). Acetone was used to elute the residual yellow that appeared in the packing. All eluates were collected, combined, and rotary evaporated to remove solvent. Samples were freeze-dried to obtain a concentrated powder and stored at 0°C.

Isolation of Flavonol Fractions

Sephadex LH-20 (3.0 x 22.0 cm), a hydroxypropylated, cross-linked dextran resin with an exclusion limit of 4-5 kDa and flow rate capacity of < 60 cm/hr, was used to isolate flavonoid derivatives. A glass column is packed with glass wool and sand. The Sephadex resin (40 g) was pretreated by swelling in 70% MeOH for 3 hours at room temperature. Particles were decanted, resuspended, and poured into the column, which was allowed to stand overnight in 70% MeOH before separation. Concentrated fruit extract (1 g) was dissolved in a minimal volume of formic acid/water/methanol (1:50:48.9 v/v/v) and loaded onto the column. After 15 minutes of absorption, phenolic compounds were eluted using 70% MeOH, and colored bands were collected separately. Brown bands containing proanthocyanidins were eluted with 70% acetone, followed by 100% acetone to elute any remaining proanthocyanidins. All collected fractions were rotary evaporated, frozen, lyophilized, and stored for further analysis.

Identification of Compounds by HPLC

Chromatographic Conditions

High-performance liquid chromatography (HPLC) analysis was performed using an Atlantis T3 C18 column (4.6 x 150 mm, 3 µm) on a Waters HPLC system equipped with a pump (Empower e2695), a photodiode array (PDA) detector (Waters 2998), an online degasser, and an autosampler. Mobile phase A consisted of water/phosphoric acid (99.5:0.5, v/v), and mobile phase B contained water/acetonitrile/glacial acetic acid/phosphoric acid (50:48.5:1.0:0.5, v/v/v/v). A reverse-phase gradient program was used, developed by Liang Xue 2021 and modified by Maureen Otieno (2023). The injection volume was 20 µL, with a flow rate of 0.900 mL/min and a total run time of 31 minutes. Chromatographs were recorded at 520, 355, 310, 280 nm to measure anthocyanidins, flavonols, phenolic acids, and proanthocyanidins, respectively.

Standard Preparation 

A stock solution of the standard was prepared by dissolving 2 mg of each standard in 2 mL of HPLC-grade methanol to obtain a concentration of 1000 ppm. Serial dilutions were performed to obtain concentrations of 100, 50, 25, 12.5, 6.25, and 3.13 ppm. Solutions were sonicated and filtered through a 0.45 mm syringe filter before injecting 20 µL into the HPLC system.

Sample Preparation

Extract samples (2 mg/5 mg/10 mg) were dissolved in 1 mL of methanol to obtain the desired concentration. Solutions were sonicated and filtered through a 0.45 mm PTFE syringe filter and 20 µL of each sample was injected into the HPLC system.

Folin-Ciocalteau Assay

The quantification of total polyphenolic content in both the crude fruit and pomace extracts was measured using a modified protocol of the Folin-Ciocalteau assay as described by Ainsworth et al.13. 20 mg of cranberry/cranberry pomace powder samples were weighed in triplicate and dissolved in 2 mL of ice-cold 95% v/v methanol to obtain a concentration of 10 mg/mL. Samples were ultrasonicated for 30 minutes and refrigerated overnight. The following day, samples were centrifuged for 5 minutes at 10,250 rpm. 40 µL of the supernatant was collected into duplicate Eppendorf tubes and reacted with 800 µL of 10% v/v F-C reagent. The solution was vortexed and allowed to stand for five minutes. 800 µL of sodium carbonate was then pipetted, followed by 360 µL of distilled water. A blank sample was prepared by substituting the stock solution with 40 µL of 95% v/v methanol. Samples were incubated in the dark at room temperature for two hours before measuring absorbance. 200 µL of respective solutions were pipetted in triplicate onto a 96-well microplate. Absorbance was measured at 760 nm using a SpectraMax 190 microplate reader. Total phenolic content was calculated as gallic acid equivalents (GAE) using the regression equation between gallic acid standards at A760 nm, using the following calculation:

T = C/CX MW

T = total phenolic content in mg/g, in GAE (gallic acid equivalents)

C = concentration of gallic acid established from the calibration curve in mg/m

C1 = concentration of the extract in mg/mL

MW = the molecular weight of gallic acid

DPPH Radical Scavenging Assay

The DPPH radical scavenging activity of both quercetin and quercetin-3-galactoside fractions was evaluated according to a modified method of Baliyan et al.14. 5 mg of each respective fraction was dissolved in 20 mL of methanol, vortexed briefly, and subjected to ultrasonication to ensure samples were fully dissolved and obtain a concentration of 250 µg/mL. From this stock solution, a serial dilution was performed to obtain the following range of concentrations: 250, 125, 62.5, 31.25, 15.63, 7.81, 3.91, 1.95 µg/mL. A DPPH solution (300 µM)was prepared by dissolving 5 mg of 2,2-diphenyl-1-picrylhydrazyl in 50 mL of methanol. 100 µL of each respective concentration was pipetted in triplicate onto a 96-well microplate. 25 µL of DPPH solution was added and mixed by pipetting. A control solution was prepared by pipetting 100 µL of methanol with 25 µL of DPPH. The blank solution contained 125 µL of methanol. The plate was incubated in the dark at room temperature for 30 minutes, and absorbance was measured at 517 nm using a SpectraMax 190 microplate reader. The following formula was used to compute percent inhibition:

DPPH % Inhibition = A– (A1 -Ab)/A0 x 100%

A0: Control absorbance with no radical scavenger (DPPH + MeOH)

A1: Sample absorbance (DPPH + scavenger)

Ab: Blank absorbance (MeOH)

Ferric Reducing Antioxidant Power

The ferric reducing power of quercetin and quercetin-3-galactoside fractions was measured using a modified version of the method described by Gashaye et al.15. Samples were prepared by dissolving 5 mg of extract in 20 mL of methanol, briefly vortexing, followed by ultrasonicating to obtain a concentration of 250 µg/mL. From this stock solution, a serial dilution was performed to obtain the following range of concentrations: 250, 125, 62.5, 31.25, 15.63, 7.81, 3.91, 1.95 µg/mL. A FRAP working solution was prepared containing 25 mL of acetate buffer (300 mM, pH 3.6), 2.5 mL of 10 mM TPTZ in 340 mM HCl, and 2.5 mL of 20 mM FeCl3 • 6H2O solution (10:1:1 v/v/v). This working solution was designated for usage within three hours of preparation. 20 µL of the respective sample was pipetted in triplicate into a 96–well microplate, followed by the addition of 180 µL of FRAP working solution and thoroughly mixed by pipetting. As a control, 20 µL of methanol and 180 µL of FRAP working solution were plated in triplicate. A blank solution consisting of 200 µL of methanol was plated in triplicate to account for background absorbance. The plate was incubated at room temperature in the dark for 30 minutes. Absorbance was measured at 593 nm using a SpectraMax 190 microplate reader. Results were calculated as percent reducing power using the following formula:

FRAP % = A1 -Ab/Ax 100%

A1: Sample absorbance (FRAP + reducer)

Ab: Blank absorbance (MeOH)

ABTS Radical Scavenging Assay

The relative ability of the flavonol fractions to scavenge ABTS was determined using a colorimetric assay as described by Tomasina et al16. A 7 mM ABTS stock solution was prepared by dissolving 0.1921 g of ABTS in 50 mL phosphate-buffered saline (PBS). A 2.45 mM potassium persulfate solution was prepared by dissolving 0.0331 g of potassium persulfate in 50 mL PBS. Equal aliquots (50 mL) of 7 mM ABTS and 50 mL 2.45 mM potassium persulfate solution were mixed and allowed to stand in the dark at room temperature for 16 hours to generate the ABTS•+ radical cation. The resulting ABTS•+ solution was diluted with 0.01 M PBTS to an absorbance of 0.70 ± 0.02 at 734 nm. Once adjusted, the solution was left to stabilize for 30 minutes and monitored for significant changes in absorbance. The reaction was initiated by reacting 190 µL of diluted ABTS•+ reagent with 10 µL of extract in triplicate wells of a 96-well microplate. A control was prepared using 10 µL of methanol in addition to 190 µL of ABTS•+, with 200 µL of methanol serving as the blank. Absorbance was read at 734 nm using a SpectraMax190 microplate reader. The following formula was used to compute percent inhibition:

% Inhibition = A– (A1 -Ab)/A0 x 100%

A0: Control absorbance with no radical scavenger (ABTS•+ + MeOH)

A1: Sample absorbance (ABTS•+ + scavenger)

Ab: Blank absorbance (MeOH)

Acetylcholinesterase In Vitro Enzyme Assay

The AChE inhibitory activities of the isolated fractions were examined using Ellman’s method18 in conjunction with the AmpliteTM Colorimetric Acetylcholinesterase Assay Kit (AAT Bioquest, Inc.), and optimized based on the procedure described by Koseki et al17. One milligram of extract was dissolved in 1 mL of DI water with 0.4% DMSO. A series dilution was performed to obtain the following concentrations: 250, 125, 62.5, 31.25, 15.63, 7.81, 3.91, 1.95 µg/mL. 10 µL of the sample solutions were pipetted into triplicate onto a 96-well microplate, followed by 30 µL of assay buffer and 10 µL of acetylthiocholinesterase (200 mU/mL in assay buffer). Galantamine (0.1 mg/mL), a selective inhibitor of AChE, was used as a positive control. Samples were incubated in the dark at room temperature for 20 minutes to facilitate binding. The reaction was initiated by adding the working reagent containing 5,5’dithiobis(2-nitrobenzoic acid) and acetylthiocholine iodide, which, upon hydrolysis, forms thiocholine. Thiocoline reacts with the reagent to form a yellow 5-thio-2-nitrobenzoate anion. After 30 minutes of incubation, absorbance was measured at 405 nm using the SpectraMax 190 microplate. The AChE inhibitory activity was calculated as follows19:

Percent Inhibition = E x S/E x 100%

E: Activity of enzyme without the inhibitor

S- Sample Absorbance

*E and S were each subtracted by their respective blank

Results and Discussions 

HPLC Quantification of Polyphenolic Compounds in MQ & CP

Polyphenolic constituents were characterized using reversed-phase high-performance liquid chromatography (HPLC) to compare both Mullica Queen (MQ) and cranberry pomace (CP) extracts for their polyphenolic content at various wavelengths. Anthocyanins. flavonols, phenolic acids, and proanthocyanidin chromatograms were extracted at 520, 355, 310, and 280 nm, respectively, and compared to standard compounds for identification. The detection of these polyphenolic compounds is relevant to their bioactive potential and provides insight into their active antioxidant components. Compounds marked in orange were identified using a published reference20 or require standard spiking. Relevant peak wavelengths are detailed below the chromatograms.

Fig. 3. HPLC Chromatogram of Mullica Queen (MQ) Fruit Extract (Run 3, 5 mg/mL) at 520 nm

Fig. 4. HPLC Chromatogram of Cranberry Pomace (CP) Extract (Run 3, 10 mg/mL) at 520 nm

HPLC chromatograms were extracted at 355 nm to reveal flavonoids present in both MQ and CP extracts, with several peaks corresponding to known flavonol standards. Peaks observed at retention times 11.79 min and 13.16 min in the MQ fruit extract were identified as myricetin-3-O-galactoside and quercetin-3-O-galactoside, respectively (Fig. 5). The aglycones of these compounds, myricetin and quercetin, were confirmed to be present in the CP extract at later elution times of approximately 18.07 min and 22.21 min, respectively (Fig. 6).

Fig. 5. HPLC Chromatogram of Mullica Queen (MQ) Fruit Extract (Run 3, 5 mg/mL) at 355 nm

Although the cranberry pomace was run at a higher concentration, the MQ extract demonstrates a broader distribution and abundance of flavonols. The difference in composition contributes to their antioxidant performance. The profiling of the whole extracts confirms the presence of the target quercetin aglycone and its glycoside, but also provides information as to other phytochemicals that may contribute to their observed antioxidant effect.

Fig. 6. HPLC Chromatogram of Cranberry Pomace (CP) Extract (Run 3, 10 mg/mL) at 355 nm

Fig. 7. HPLC Chromatogram of Mullica Queen (MQ) Fruit Extract (Run 3, 5 mg/mL) at 310 nm

Fig. 8. HPLC Chromatogram of Cranberry Pomace (CP) Extract (Run 3, 10 mg/mL) at 310 nm

Fig. 9. HPLC Chromatogram of Mullica Queen (MQ) Fruit Extract (Run 3, 5 mg/mL) at 280 nm

Fig. 10. HPLC Chromatogram of Cranberry Pomace (CP) Extract (Run 3, 10 mg/mL) at 280 nm

Quantification of Polyphenolic Compounds in MQ Extract

Quantitative analysis of the MQ extract reveals that quercetin-3-O-galactoside was the most abundant polyphenolic compound with an average concentration of 13.9 ± 0.19 mg/g of desugared extract. Quercetin glycosides constituted much of the polyphenolic content in this extract. Although glycoside derivatives are present in high concentrations in this sample, trace amounts of free quercetin were also detected, though present at significantly lower concentrations. CP extract likely contains higher levels of quercetin aglycones, potentially due to processing conditions during juice extraction, where these water-insoluble aglycones may be concentrated in the pomace.

Table 1. Mullica Queen Polyphenolic Content Quantification

Compound  Average Concentration (mg analyte/g of Desugared Extract) 

FLAVONOLS 

Quercetin  0.33 ± 0.00 
Quercetin-3-O-Galactoside  13.9 ± 0.19 
Myricetin-3-O-Galactoside  12.2 ± 0.09 
Isorhamnetin  0.11 ± 0.01 

PHENOLIC ACIDS 

p-Coumaric Acid  2.55 ± 0.02 
Ideain Chloride  0.51 ± 0.0 
Peonoidin-3-O-Galactoside Chloride  6.65 ± 0.04 

ANTHOCYANINS 

Cyanidin-3-O-Galactoside  0.51 ± 0.0 
Peonidin-3-O-Galactoside  0.60 ± 0.0 

PROANTHOCYANINS 

PACs A  7.39 ± 0.02 
PACs B  3.5 ± 0.35 

 Isolation and Confirmation of Quercetin & Quercetin-3-galactoside in Isolated Fractions

Fractionation of the fruit and pomace extracts using Sephadex gel filtration chromatography was proven successful in the isolation of both quercetin and its glycoside (Table 2). Sephadex fractionation relies on size exclusion principles, where larger molecules elute first and smaller molecules are retained longer in the porous matrix. A peach-colored eluate corresponding to 70%methanol Eluate IV was the final fraction eluted with this solvent and confirmed to contain quercetin-3-galactoside in this MQ fraction. The compound’s strong retention onto the column indicates strong interactions with the Sephadex matrix. HPLC spectral data were used for identification as the compound strongly absorbed at 355.3 nm, corresponding to standard data and retention times (Fig. 11).

Table 2. Mullica Queen Sephadex LH-20 Fractionation Output

 

Fig. 11. MQ 70% Methanol Eluate IV Fraction Chromatogram (Run 3, 2 mg/mL) at 355 nm

A yellow-colored eluate corresponding to 70% Acetone Eluate III was confirmed to contain the aglycone quercetin as the main active component in this CP fraction. Spectral data showed a strong absorbance at 363.6 nm corresponding to the standard quercetin wavelength and retention times (Fig. 12). Solubility differences between quercetin and its more polar glycoside impact their interaction with the nonpolar stationary phase. Polar mobile phase solvent, 70% methanol, facilitates the elution of the more polar glycoside, whereas 70% acetone is strong enough and less polar to elute the aglycone with reduced polarity.

INSERT FIGURE 12 HERE

Fig. 12. CP 70% Acetone Eluate III Fraction Chromatogram (Run 3, 10 mg/mL) at 355 nm

Determination of Total Phenolic Content 

Phenolic compounds are critical for antioxidant defense to neutralize free radicals. The Folin-Ciocalteau (F-C) assay is a widely used method to measure total phenolic content (TPC). Reaction of phenolic compounds reveals their antioxidant power based on the reduction of yellow phosphotungstate-phosphomolybdate complex by antioxidants that reduce the complex to a blue chromagen measured at 760 nm21. Reducing capacities of the powdered CP and MQ extracts were measured to compare the broader polyphenolic composition of the extracts the quercetin and its glycosides were isolated. The total phenolic content was measured in milligrams of gallic acid per gram of dry powder. The MQ fruit extract exhibited a higher total phenolic content compared to the CP extract, which is consistent with the presence of more diverse polyphenolic compounds, suggesting the fruit may serve as a richer source of flavonol glycosides. The TPC of the MQ powder was significantly higher (192.1 ± 1.53 mg/g ) than CP (137.3 ± 2.28).

Fig. 13. Total Phenolic Content in CP & MQ

DPPH Free Radical Scavenging Antioxidant Activity

Another commonly used bioanalytical method to measure antioxidant activity is the 1,1-diphenyl-2 picrylhydrazyl (DPPH) assay. This method evaluates antioxidant capacity based on spectrophotometric measurements of antioxidants’ ability to scavenge DPPH free radicals. The underlying mechanism involves the reaction of DPPH· radicals with hydrogen-donating antioxidants to form a reduced hydrazine compound (DPPH-H). Upon the formation of hydrazine, a neutralized, yellow-colored solution appears, indicating the radical neutralization from its original dark purple complex, spectrophotometrically observed at 517 nm22. This assay was employed to compare the antioxidant capacity of de-sugared CP and MQ extracts, as well as their respective quercetin and quercetin glycoside fractions.

Fig. 14. (a) Percent Inhibition of DPPH radical by CP and MQ de-sugared extracts. (b) Percent inhibition of DPPH radical by CP Quercetin, and MQ Quercetin-3-Galactoside fractions

Results indicate that the MQ extract exhibited a higher percentage of DPPH radical inhibition than the CP extract, though data did not significantly vary (Fig. 14a). Both extracts displayed dose-dependent behavior with increasing concentrations resulting in greater inhibition of DPPH radicals. Analysis of the CP quercetin and MQ quercetin-3-galactoside fractions revealed significant differences in antioxidant activity at lower concentrations, as higher concentrations were omitted to demonstrate a more pronounced difference. CP-derived quercetin demonstrated higher antioxidant activity with respect to its glycoside fraction, despite originating from the CP extract with overall lower antioxidant activity and TPC. At 31.3 μg/mL, CP quercetin exhibited 81.0 ± 12.1% inhibition, outperforming its glycoside, which only reached 73.7 ± 3.79 % at the same concentration (Fig. 14b).

Ferric Reducing Antioxidant Power

Further antioxidant analysis was performed using the Ferric Reducing Antioxidant Power assay(FRAP), which measures antioxidant reduction potency based on the reduction of a colorless Fe3+- TPTZ complex into an intense blue Fe2+-TPTZ in the presence of antioxidants. This reduction in acidic medium is measured spectrophotometrically at 593 nm. Notably, the ferricreducing antioxidant power between the desugared CP and MQ extracts demonstrates differing results in respect to the DPPH assay.

Fig. 15. (a) Ferric reducing antioxidant power by CP and MQ de-sugared extracts (b) Ferric-reducing antioxidant power by CP Quercetin and MQ Quercetin-3-Galactoside fractions

Results show CP extract exhibited greater ferric reducing antioxidant power at varying concentrations (250 – 3.91 μg/mL). Although it’s important to note that the dose-dependent response is not as significant for CP extracts as previously noted, considering lower concentrations still exhibited significant reduction potential with respect to MQ (Fig. 15a). Analysis of the fractions reveals CP quercetin demonstrates significantly higher ferric reducing antioxidant power with respect to its MQ quercetin-3-galactoside fraction. At 31.3 μg/mL, CP quercetin exhibited 57.4 ± 4.29% reducing power, outperforming its glycoside, which only reached 43.5 ±1.28 % at the same concentration (Fig. 15b).

ABTS Radical Scavenging Activity

To further validate the antioxidant potential of the isolated fractions, the 2,2-azino-bis-3-ethylbenzothiazoline-6-sulphonic acid (ABTS)assay was utilized. The principle of this method relies on the electron transfer of antioxidants in reducing ABTS radical cation (ABTS•+) to its colorless, neutralized form with a decreased absorbance measured spectrophotometrically at 734 nm. Both CP quercetin and MQ quercetin-3-galactoside fractions demonstrated strong free radical scavenging at 250 μg/mL, with ABTS inhibition values of 98.2 ± 0.96 % and 95.9 ±4.88 %, respectively. CP quercetin maintained relatively higher activity, exhibiting 26.3 ±3.51% inhibition, outperforming its glycoside fraction, which showed lower inhibition at 24.9 ±2.39%. Findings from the ABTS assay, along with FRAP and DPPH results, support the conclusion that CP-derived quercetin exhibits stronger antioxidant activity than its glycosylated counterpart.

Fig. 16. Comparison of ABTS scavenging activity between MQ Quercetin-3-galactoside and CP Quercetin fractions

Inhibitory Effect on Acetylcholinesterase(AChE)

Acetylcholinesterase (AChE) is a neurodegenerative-associated enzyme that catalyzes the hydrolysis of acetylcholine, a neurotransmitter critical to memory, attention, and learning8. AChE inhibitors are used to treat Alzheimer’s disease by inhibiting the breakdown of acetylcholine. To assess their potential neuroprotective properties of cranberry and cranberry pomace-derived fractions, AChE-inhibitory activity was evaluated in vitro. This assay employed Ellman’s reagent to quantify thiocholine, a product of acetythiocholine hydrolysis by AChE. Thiocholine reacts with DTNB (5,5’-dithiobis (2-nitrobenzoic acid)) to produce a yellow complex measured at 405 nm.

Fig. 17. Inhibition of AChE activity by CP Quercetin

Using the AmpliteTM Colorimetric Acetylcholinesterase Assay Kit (AAT Bioquest, Inc.) with protocol adjustments based on a method described by Koseki et al18, CP-derived quercetin exhibited significant inhibitory activity. At a concentration of 250 μg/mL, this fraction achieved 59.0 ± 2.41% inhibition of AChE activity, while at a lower concentration, 31.3 μg/mL, the sample reached 8.85 ± 2.86% AchE inhibition. Galantamine, a clinically approved AChE inhibitor used in the treatment of dementia, served as the positive control. As subsequent trials were conducted, tests at higher concentrations with the CP quercetin extract and the evaluation of the inhibitory potential of the MQ quercetin-3-galactoside fraction demonstrated inconsistent results. A reduction in galantamine’s inhibitory effect was also observed during these trials, indicating partial degradation of the enzyme. Solubility issues also arose when dissolving the glycoside fraction, which required increased DMSO concentrations that could have impacted its inhibitory effects. Overall, while CP-derived quercetin shows promising inhibitory activity against AChE, this method requires further optimization to allow for accurate comparisons of fraction AChE inhibitory activity.

Conclusions 

The results of this study suggest that cranberry pomace-derived quercetin exhibits stronger antioxidant activity than its glycosylated counterpart, quercetin-3-galactoside, isolated from Mullica Queen (MQ) fruit extract. HPLC profiling and fractionation confirmed these compounds as the primary active constituents of their respective extracts. Despite the CP powder having lower TPC than MQ, CP-derived quercetin consistently outperformed its glycoside-rich fraction in antioxidant assays (DPPH, FRAP, ABTS) and demonstrated moderate acetylcholinesterase(AChE) inhibitory activity. It should be noted that the DPPH antioxidant activity of the CP and MQ concentrated extracts themselves is relatively statistically equivalent, with the CP extract exhibiting greater ferric reducing antioxidant power maintained throughout lower concentrations. ABTS results show relatively similar potential in reducing ABTS radical cation, although CPquercetin still demonstrated higher inhibition than its glycoside. Results highlight CPquercetin’s neuroprotective potential through its ability to scavenge reactive oxygen species(ROS). The differences in activity between the aglycone and its glycoside are likely due to the absence of the sugar moiety. Findings from this study support the conclusion that glycosylation may reduce bioactivity. Future studies will explore their effects on additional neurodegenerative-related enzymes such as monoamine oxidases (MAOs) and butyrylcholinesterase, as well as potentially conducting in vivo studies to evaluate how glycosylation impacts absorption and therapeutic potential.

Acknowledgements

This work was supported by funding from the Office of Undergraduate Research at the University of Massachusetts Dartmouth and the UMass Cranberry Health Research Center. The author also gratefully acknowledges the support of Dr. Neto, Maureen Otieno, and the rest of the Neto lab research group.

References

(1) World Health Organization. (2007, February 27). Neurological disorders affect millions globally: WHO report. Www.who.int. https://www.who.int/news/item/27-02-2007-neurological-disorders-affect-millions-globally-who-report

(2) Arias-Sánchez, R. A., Torner, L., & Fenton Navarro, B. (2023). Polyphenols and Neurodegenerative Diseases: Potential Effects and Mechanisms of Neuroprotection.Molecules (Basel, Switzerland), 28(14), 5415. https://doi.org/10.3390/molecules28145415

(3) Ayaz, M., Sadiq, A., Junaid, M., Ullah, F., Ovais, M., Ullah, I., Ahmed, J., &Shahid, M. (2019). Flavonoids as Prospective Neuroprotectants and Their Therapeutic Potential in Aging Associated Neurological Disorders. Frontiers in aging neuroscience, 11, 155. https://doi.org/10.3389/fnagi.2019.00155

(4) Chiang, M. C., Tsai, T. Y., & Wang, C. J. (2023). The Potential Benefits of Quercetin for Brain Health: A Review of Anti-Inflammatory and Neuroprotective Mechanisms. International journal of molecular sciences, 24(7), 6328.

(5) Blumberg, J. B., Camesano, T. A., Cassidy, A., Kris-Etherton, P., Howell, A., Manach, C., Ostertag, L. M., Sies, H., Skulas-Ray, A., & Vita, J. A. (2013). Cranberries and Their Bioactive Constituents in Human Health. Advances in Nutrition, 4(6), 618–632. https://doi.org/10.3945/an.113.004473

(6) Kaushik, A., Chauhan, K., & Singh, S. (2023). Neuroprotective potential of quercetin as a nutraceutical targeting fused neuroinflammation in neurological disease (pp. 623–637). Academic Press. https://doi.org/10.1016/B978-0-323-90052-2.00029-9

(7) Xie, L., Deng, Z., Zhang, J., Dong, H., Wang, W., Xing, B., & Liu, X. (2022).Comparison of Flavonoid O-Glycoside, C-Glycoside, and Their Aglycones on Antioxidant Capacity and Metabolism during In Vitro Digestion and In Vivo. Foods (Basel, Switzerland), 11(6), 882.

8) Cleveland Clinic. (2022, December 30). Acetylcholine (ACh). Cleveland Clinic.

(9) ArborAssays. (2018, February 8). Cholinesterases: Neurotransmitter Control Systems – Arbor Assays. Arbor Assays.

(10) Chen, K., & Shih, J. C. (1997). Monoamine Oxidase A and B: Structure, Function, and Behavior. Advances in Pharmacology, 42(1054-3589), 292–296. ScienceDirect.

(11) Mayo Clinic. (2024, September 27). Parkinson’s Disease. Mayo Clinic.

(12) Román, S., Sánchez-Siles, L. M., & Siegrist, M. (2017). The importance of food naturalness for consumers: Results of a systematic review. Trends in Food Science &Technology, 67(0924-2244), 44–57. https://doi.org/10.1016/j.tifs.2017.06.010

(13) Ainsworth, E. A., & Gillespie, K. M. (2007). Estimation of total phenolic content and other oxidation substrates in plant tissues using Folin-Ciocalteu reagent. Natureprotocols, 2(4), 875–877. https://doi.org/10.1038/nprot.2007.102

(14) Baliyan, S., Mukherjee, R., Priyadarshini, A., Vibhuti, A., Gupta, A., Pandey, R.P., & Chang, C. M. (2022). Determination of Antioxidants by DPPH Radical Scavenging Activity and Quantitative Phytochemical Analysis of Ficus religiosa. Molecules (Basel, Switzerland), 27(4), 1326. https://doi.org/10.3390/molecules27041326

(15) Gashaye, M.B., Birhan, Y.S. Phytochemical constituents, antioxidant and antibacterial activities of Plectocephalus varians (A. Rich.) C. Jeffrey ex Cufod root extracts. BMC Complement Med Ther 23, 135 (2023). https://doi.org/10.1186/s12906-023-03919-8

(16) Tomasina, F., Carabio, C., Celano, L., & Thomson, L. (2012). Analysis of two methods to evaluate antioxidants. Biochemistry and molecular biology education: abimonthly publication of the International Union of Biochemistry and Molecular Biology, 40(4), 266–270. https://doi.org/10.1002/bmb.20617

(17) Ellman G, Courteny K, Anderes V, Featherstone R (1961) A new and rapid colorimetric determination of acetylcholinesterase activity. Biochem Pharmacol 7:88–95

(18) Koseki, Y., Nishimura, H., Asano, R., Aoki, K., Shiyu, L., Sugiyama, R., &Yamazaki, M. (2025). Isolation of new indole alkaloid triglucoside from the aqueous extract of Uncaria rhynchophylla. Journal of natural medicines, 79(1), 28–35

(19) Samani, P., Costa, S., & Cai, S. (2023). Neuroprotective Effects of Blueberries through Inhibition on Cholinesterase, Tyrosinase, Cyclooxygenase-2, and Amyloidogenesis. Nutraceuticals, 3(1), 39-57.

(20) Wu, X., Xue, L., Tata, A., Song, M., Neto, C. C., & Xiao, H. (2020). Bioactive Components of Polyphenol-Rich and Non-Polyphenol-Rich Cranberry Fruit Extracts and Their Chemopreventive Effects on Colitis-Associated Colon Cancer. Journal of agricultural and food chemistry, 68(25), 6845–6853. https://doi.org/10.1021/acs.jafc.0c02604

(21) Pérez, M., Dominguez-López, I., & Lamuela-Raventós, R. M. (2023). The Chemistry Behind the Folin-Ciocalteu Method for the Estimation of (Poly)phenol Content in Food: Total Phenolic Intake in a Mediterranean Dietary Pattern. Journal of agriculturaland food chemistry, 71(46), 17543–17553

Research in Bioengineering

Formulation of Smart Hydrogels with Embedded Nanoparticles for Targeted Melanoma Treatment 

By: Emma Thiboutot
Background Information 

The problem discussed within this proposed project is melanoma skin cancer, with a focus on its impact and the current treatment options available. Every year, there are approximately 5.4 million cases of skin cancer diagnosed in the United States alone. This makes it the most common type of cancer (Skin Cancer Foundation, 2023). The default treatments for melanoma include, but are not limited to, chemotherapy, radiation, and surgical interventions. All of which are invasive and often significantly burden patients. These treatments are not limited to affecting the patient physically, but they can also impose financial strains during and after recovery, ultimately impacting the patient’s quality of life negatively beyond their diagnosis. Furthermore, the importance of addressing this problem goes beyond individual cases, but also affects families, the healthcare system, and society at large. Melanoma treatments are very expensive, and they can emotionally, physically, and financially drain the patients and their families. An effective treatment that manages skin cancer requires improvements in public health care policies, which would help to reduce the prevalence of skin cancer, enhance the current care options, and ensure equal access to prevention and treatment options. There are current innovations for the treatment of skin cancer that have included the development of hydrogels for localized and controlled drug delivery. This helps in minimizing side effects and improving the efficiency of the drugs that are delivered at the tumor site. There have been studies conducted by Ye et al. (2018) and Zheng et al. (2015) that have highlighted the use of injectables and pH-responsive hydrogels that have been shown to directly deliver multiple anti-cancer agents to affected tumors and sites. This would enhance therapeutic efficiency while sparing healthy tissues from the harmful effects of anti-cancer drugs. Additionally, there has been research conducted by Elhabal et al. (2024) and Ahmed et al. (2023) to support using heat-responsive and ROS-responsive hydrogels for localized cancer therapy. These studies highlight further advancements in responsive and targeted drug delivery systems. Although all these studies show research promise in optimizing anti-cancer drug delivery systems, there are challenges that still exist regarding biocompatibility and production scaling.

Research Question(s)/Thesis Statement/Aims

The proposed solution to this problem is to develop a novel hydrogel system loaded with an anti-cancer drug, doxorubicin, and an NIR responsive dye, indocyanine green, to treat melanoma skin cancer non-invasively with SMART-activated drug-releasing hydrogels. The project’s goal is to prove that these hydrogels can effectively target and eliminate melanoma cells with minimal impact on surrounding healthy tissues. The specific question that is being researched throughout this question is: Can PVA-chitosan hydrogel encapsulating doxorubicin and light-sensitive compounds demonstrate a significant cytotoxicity against melanoma cells? The research will focus on creating a PVA-chitosan crosslinked hydrogel loaded with an anti-cancer drug and SMART compounds.

Objectives of the Proposed Project

The completion of this project will contribute significantly to the field of skin oncology by providing a more targeted and potentially more effective treatment for melanoma with reduced side effects. Furthermore, this research could provide a foundation for future studies to explore the use of drug delivery in hydrogel formulations for various types of cancer. It could also influence the design of multifunctional biomaterials with engineered delivery systems.

Description of the Proposed Work

This project is an in vitro study, meaning it does not involve human participants but instead uses A375 cell lines, a type of skin cancer cell. These cells will be sourced from Dr. Tracie Ferreira at the University of Massachusetts–Dartmouth’s Bioengineering Department. Additional cells required will be purchased from Sigma-Aldrich. The research will be experimental, conducted in a controlled laboratory setting to test the effectiveness of different hydrogel formulations on melanoma cells. The experimental design for this project will include four treatment groups. First, there will be the formulated hydrogel without any loading, which will be the control. Then, there will be the formulated hydrogel loaded with indocyanine green for its activation due to near-infrared light. Then, there will be the hydrogel loaded with doxorubicin, a chemotherapy drug. Finally, the last formulation of the hydrogel will be the combination of doxorubicin and indocyanine green to offer a controlled drug release at the affected site. The main focus of this study is to observe the viability of melanoma cells after treatment with hydrogels. Cell viability, the dependent variable, will be assessed through the cell viability assay Alamar Blue, which measures metabolic activity, and live/dead cell assays. These tests help quantify data by measuring changes in fluorescence, which are then scored as a percentage of viability compared to untreated control groups. The research will unfold in two phases: 1. In the first phase, we will create light-responsive PVA-chitosan hydrogels containing the doxorubicin and/or the indocyanine green. We will evaluate the physical and chemical properties of these hydrogels, such as drug encapsulation efficiency and release kinetics, using several techniques: Scanning Electron Microscopy (SEM) to visualize the surface structure of the hydrogels, Thermogravimetric Analysis (TGA) to assess the thermal stability, Differential Scanning Calorimetry (DSC) to study the thermal properties, and Fourier-Transform Infrared Spectroscopy (FTIR) to identify the chemical composition. These characterizations will measure properties like light responsiveness, color changes, and drug delivery times. 2. In the second phase, we will test these hydrogels in vitro using A375 melanoma cell cultures in cell culture inserts. The response of these cells to the hydrogel treatment will be monitored by measuring cell viability using Alamar Blue and a live/dead cell assay conducted through Confocal Microscopy. Data collection will be electronic, ensuring accurate, real-time capture during experiments, and analyzed using software like GraphPad Prism.

Clear Statement of the Originality of the Proposed Work

This research is original in its approach by incorporating cranberry extract, a natural substance, into hydrogel formulations for treating melanoma, an application not extensively explored in current literature, and comparing its efficacy to a chemotherapy drug, doxorubicin, with the same hydrogel delivery system. This novel approach could provide significant insights into the potential of natural compounds in cancer treatment, challenging traditional methods and potentially setting a precedent for future therapeutic developments. All laboratory procedures will adhere to standardized protocols for the ethical handling and treatment of cell cultures, data recording, and analysis.

Materials and Methods
Hydrogel Preparation

The PVA-chitosan hydrogels were prepared using a freeze-thaw and genipin crosslinking method to utilize both physical and chemical crosslinking methods. Medium molecular weight chitosan was dissolved in 9 mL of 2% (v/v). lactic acid, and subsequently 3 mL of 10% (w/v) poly(vinyl alcohol) (PVA) solution, prepared in distilled water, was added under continuous stirring. For crosslinking, 18 μL of genipin was incorporated into the mixture. Formulations were created into four groups: CP6 (control), CP6-DOX (loaded with 10 mg doxorubicin), CP6-ICG (loaded with 10 mg of indocyanine green), and CP6-DOX-ICG (co-loaded with 10 mg each of DOX and ICG). Drugs were added during the mixture stage and before the freeze-thaw cycles. The mixtures were cast into petri dishes and subjected to three freeze-thaw cycles (freezing at -20 °C for 18 hours and followed by 6 hours of thawing at room temperature). Additionally, the hydrogels were allowed to crosslink at room temperature on the bench for an additional 3 days.

Freeze-drying and Sample Preparation

All hydrogel samples were freeze-dried for 12 hours before thermal and morphological characterization tests. Freeze-dried samples were used directly for SEM imaging without sputter coating, due to equipment limitations.

Tensile Strength Testing

Mechanical testing was performed using a Shimadzu tensile tester to assess the tensile strength of each hydrogel formulation. The samples were tested after crosslinking.

Thermal Analysis

Thermogravimetric analysis and differential scanning calorimetry were done to evaluate the thermal stability and phase transitions of the hydrogel formulations. Differential Scanning Calorimetry (DSC) analysis was done using the Thermal Analyzer DC25. An average of 7 mg samples were used for the DSC, and an average of 15 mg samples were used for the TGA after freeze drying. TGA was performed from 30 °C to 500 °C under a nitrogen atmosphere, while DSC was carried out from 30 °C to 400 °C to capture water loss and the polymer transitions.

Scanning Electron Microscopy

The surface and cross-section areas of the hydrogels were analyzed using a HITACHI 3700nVP scanning electron microscope (SEM) at an accelerating voltage of 2 kV. Samples were mounted onto the stubs without sputter coating and imaged at magnifications x50, x150, and x250. Cross-section samples were prepared by cutting the freeze-dried hydrogels and positioning them vertically.

Swell Test

The swelling behavior of the hydrogel formulations CP6, CP6-ICG, and CP6-DOX was evaluated in three different solutions: ultra-pure water, Dulbecco’s Modified Eagle Medium (DMEM), and phosphate-buffered saline (PBS, pH 7.4). Freeze-dried hydrogel samples were individually weighed to obtain an initial dry weight and then submerged in 5 mL of the respective medium. Samples were incubated at 37 °C to mimic physiological conditions. At time intervals of 5 min, 10 min, 1 hr, 2 hr, 4 hr, 6 hr, 8 hr, 24 hr, 30 hr, 48 hr, 72 hr, 96 hr, 120 hr, 7 days, 8 days, 9 days, and 10 days. Hydrogels were gently removed from the medium, blotted with filter paper to remove surface moisture, and weighed. Between measurements, all samples were returned to the incubator in fresh medium to continue swelling under consistent conditions.

Cell Viability Assay (alamarBlue™)

Cell viability was assessed using the alamarBlue™ assay to evaluate the cytotoxic effects of CP6 hydrogel formulations. A375 human melanoma cells were seeded in 96-well plates at a density of 25,000 cells per well and incubated at 37 °C in a humidified atmosphere with 5% CO2 for 24 hours to allow cell attachment. Treatment groups included direct contact and cell culture inserts of each of the formulations: CP6, CP6-ICG, CP6-DOX, and CP6-DOX-ICG. Additional groups were exposed to near-infrared (NIR) laser irradiation (808 nm) for photothermal activation when applicable.

For insert-based experiments, the sterile hydrogels were prepared and placed into inserts suspended above seeded cells to simulate indirect drug exposure. NIR-activated groups were exposed for 10 minutes before incubation. Alamar Blue reagent was added to each well at 24, 48, and 72 hours post-treatment. After 2 hours of incubation with the reagent, absorbance was measured at 570 nm and 600 nm using a microplate reader. Cell viability was calculated and normalized to untreated control cells and the working solution of Alamar Blue.

Results and Discussion
Tensile Strength

Tensile strength testing and data were collected using the Shimadzu machine that has been provided for use in the biomechanics laboratory. The tensile strength testing was conducted on three different formulations of the PVA-chitosan hydrogels. There was the CP formulation, the ICG incorporated CP formulation, and the NPC incorporated hydrogel, which includes the ICG and cranberry nanoparticles. The objective of this testing was to evaluate how the baseline hydrogel performs, while also evaluating how the hydrogel performs when there are added compounds into it under induced tensile load for real-time applications in the future. These tests have determined maximum and minimum force with stained, average tensile strength, and standard deviation to assess the strength and the consistency based on all different types of hydrogel formulation.

Force N
Average  0.409954
SD 0.297193
Max 1.038734
Min 0.05095

Table 1: A table representing the CP hydrogel formation tensile data, including the average, standard deviation, maximum, and minimum forces in Newton values.

 

Displacement  mm
Average 20.6691
SD 11.93479
Max 41.32924
Min 0.0021

Table 2:A table representing the CP hydrogel formation tensile data, including the average, standard deviation, maximum, and minimum displacements in millimeter values.

 

Figure 1: This is the force versus displacement curve for the CP hydrogel. This graph displays the typical nonlinear tensile behavior with minimal outliers and a linear trendline. The trendline equation of y=0.051x-0.0959 represents the best-fit linear relationship approximation of the sample’s tensile behavior.

 

Force N
Average 0.368162
SD 0.323111
Max 1.173194
Min -0.03352

Table 3: A table representing the ICG hydrogel formation tensile data, including the average, standard deviation, maximum, and minimum forces in Newton values. 

 

Displacement mm
Average 20.6691
SD 11.93479
Max 41.32924
Min 0.0021

Table 4: A table representing the ICG hydrogel formation tensile data, including the average, standard deviation, maximum, and minimum displacements in millimeter values.

 

Figure 2: This is the force versus displacement curve for the ICG hydrogel. This graph displays the typical nonlinear tensile behavior with minimal outliers and a linear trendline. The trendline equation of y=0.0258x-0.1655 represents the best fit linear relationship approximation of the sample’s tensile behavior.

 

Force mm
Average 0.301788
SD 0.228036
Max 0.798583
Min -0.04087

Table 5: A table representing the DOX hydrogel formation tensile data, including the average, standard deviation, maximum, and minimum forces in Newton values.

 

Displacement mm
Average 12.21906
SD 7.05618
Max 24.42903
Min 0.001833

Table 6: A table representing the NPC hydrogel formation tensile data, including the average, standard deviation, maximum, and minimum displacements in millimeter values.

 

Figure 3: This is the force versus displacement curve for the DOX hydrogel. This graph displays the typical nonlinear tensile behavior with minimal outliers and a linear trendline. The trendline equation of y=0.0317x-0.0851 represents the best fit linear relationship approximation of the sample’s tensile behavior.

 

CP6 (1) ICG (1) DOX (1)
Force N Force N Force N
Average 0.409954 Average 0.368162 Average 0.301788
SD 0.297193 SD 0.323111 SD 0.228036
Max 1.038734 Max 1.173194 Max 0.798583
Min -0.05095 Min -0.03352 Min -0.04087
CP6 (1) ICG (1) PCN (1)
Displacement mm Displacement mm Displacement mm
Average 9.922375 Average 20.6691 Average 12.21906
SD 5.730193 SD 11.93479 SD 7.05618
Max 19.83097 Max 41.32924 Max 24.42903
Min 0.0018 Min 0.0021 Min 0.001833

Table 7: Combined summary of tensile force and displacement data collected from all hydrogel groups: CP, ICG, and DOX. This table shows a direct comparison of the mechanical performance between the three formulations.

 

Figure 4: A bar graph comparing the average tensile force in newtons of all three hydrogel formulations, labeled respectively with a 5% error bar for each, indicating the relative variability. Hydrogel CP shows the highest average force, suggesting the strongest mechanical integrity, while DOX shows the lowest average force, suggesting the weakest mechanical integrity. Average values are listed on top of the bar graph.

 

Figure 5: A bar graph comparing the average displacement in millimeters of all three hydrogel formulations, labeled respectively with a 5% error bar for each, indicating the relative variability. The ICG hydrogel shows the greatest deformation before failure, suggesting that it has the highest elasticity of all of the hydrogel formations. Contrary, hydrogel CP displayed the least elasticity. Average values are listed on top of the bar graph.

The tensile strength testing data that have been collected provide valuable insights into how the formulation modification of the PVA-chitosan hydrogel affects its mechanical integrity, specifically in terms of the tensile strength and elasticity. When comparing the hydrogel formulations CP, ICG, and DOX, there were clear trends that can be observed regarding the average force, maximum displacement, and the variability between all three hydrogels.

The CP hydrogel was observed to have an average tensile strength of 0.41 N with a relatively low standard deviation. This supports the idea that the original hydrogel formulationwith no extra additions, forms a well-structured and dense hydrogel. The CP hydrogel does have a relatively lower average displacement, which further supports its classification as a stiffer material. The CP hydrogel is concluded to be a stiffer material and the best option for structural support applications compared to the other formulations.

The ICG hydrogel that incorporates the indocyanine dye shows the highest maximum tensile force of 1.17 N, but it also has the highest variability. This suggests that although the ICG hydrogel can withstand a higher maximum tensile force, due to perhaps a localized stiffness, its overall impact is less consistent due to its lower average of 0.37 N than the CP hydrogel. Furthermore, the ICG hydrogel had the greatest displacement with a value of 20.67 mm, which indicates a higher elasticity. To point out a balance between the two measurements, the ICG hydrogel has a mechanical consistency and responsiveness, making it potentially useful for drug delivery applications in the future.

The DOX hydrogel has both ICG and cranberry nanoparticles incorporated into it. These hydrogels displayed the lowest average force of 0.3 N and a median average displacement of 12.22mm. However, it did have the lowest standard deviation, which suggests that the mechanical testing was the most consistent across all of the differently formulated hydrogels. The hypothesis is that the nanoparticles that are embedded in the hydrogel affect the crosslinking density, which, as a result, weakens the tensile strength. Alternatively, this shows the potential for the improvement of uniformity if well dispersed. This makes the DOX hydrogel the best fit for a controlled delivery system, with emphasis on when the mechanical load is minimal.

Furthermore, the trendlines that are displayed respectively to each hydrogel formation quantify the slope of the force versus displacement relationship. A steeper slope, like the CP hydrogel, corresponds to the higher stiffness, while a shallower slope, like the ICG and DOX hydrogels, corresponds to a more elastic behavior. These equations are represented in the form of y=mx+b, where the slope is the rate of increase in force with respect to the displacement, which means how much force is required to stretch the material. For the CP formulation, the equation is:

= 0.051 − 0.0959

The slope of 0.51 shows that the CP hydrogel needs a higher amount of force per millimeter of stretch, which is supported by the high tensile strength average and low displacement average, and confirms the stiffness of the CP hydrogel. In contrast, the ICG hydrogel displayed an equation of:

= 0.0258 − 0.1655

This shallower slope of 0.0258 shows that the ICG hydrogel is more elastic and displays a more stretchable nature, which requires less force per millimeter. For the DOX hydrogel that contains the ICG and cranberry nanoparticles, the equation that is displayed is:

= 0.0317 − 0.0851

The slope of 0.0317 demonstrates that this hydrogel is more deformable than CP6, but it is more resistant than the ICG hydrogel. The PCN hydrogel displays the lowest average force of 0.302 N but has a moderate displacement of 12.22 mm. However, the PCN hydrogel’s trendline displays a relatively uniform response to force loading and is supported further by its low standard deviation. It is important to note that these slope values are representative of Young’s Modulus.

It is also important to note that only one sample from each hydrogel formulation was successfully tested due to the time constraints and issues with sample integrity. Several other samples of each formulation were mechanically damaged before testing, or they could not be tested in time for the submission of this report. This limits the statistical prevalence of the findings. For the future, results should include multiple test samples of each hydrogel formulation.

Fourier Transform Infrared Analysis

Figure 6: The FTIR results are shown above, visually compared. CP6 is the unloaded hydrogel, CP6-DOX is a doxorubicin-loaded hydrogel, CP6-ICG is the indocyanine green-loaded hydrogel, and CP6-DOX-ICG is the doxorubicin and indocyanine green-loaded hydrogel. The plot visually shows absorption peaks for each formulation.

The FTIR results confirm successful crosslinking and a successful functional integration of the ICG and Doxorubicin within the CP6 base hydrogel. All of the different formulations do show a similar baseline upon further analysis, suggesting that the chemical backbone of the hydrogels remains stable throughout different variations of loading withstood.

For the CP6 hydrogel, the absorption span was observed to be around 3200-3400 cm-1. This observation corresponds to the PVA and chitosan’s hydrogen bond interactions and the presence of the hydroxyl and amine groups. The peak is around 1640 cm-1, which is attributed to double-bonded oxygen and carbon, which confirms the genipin crosslinking with chitosan’s primary amine groups. Other peaks observed range around 1050-1150 cm-1 and correspond to C-O-C stretching, indicating the polysaccharide backbone.

For the CP6-Doxorubicin hydrogel, there are observed subtle increases in peak intensities near 1600 cm-1 and 1250 cm-1, which could indicate possible hydrogen bonding between the doxorubicin and hydrogel network. For the CP6-ICG hydrogel, the integration of the ICG is observed to shift towards higher wavenumbers, which is potentially due to the dispersion of ICG ’sulfonate groups within the hydrogel.

For the CP6-DOX-ICG hydrogel, both features unique to each compound, as observed above, are present, but they do not introduce any significant peaks outside of the expected functional group regions. This would suggest that the doxorubicin and the indocyanine green are physically entrapped within the hydrogel rather than chemically bonded to the hydrogel structure.

Thermogravimetric Analysis

Figure 7: TGA analysis of the CP6 hydrogel formulations. The CP6 base hydrogel (top left), CP6-ICG (top right), CP6-DOX (bottom left), and CP6-DOX-ICG (bottom right) were subjected to the TGA to assess the thermal stability. All samples did display degradation with initial water loss and were then followed by polymer decomposition. The doxorubicin-loaded and the indocyaninegreen-loaded hydrogel exhibited lower thermal stability compared to the base CP6 hydrogel, while CP6-DOX-ICG showed the greatest total weight loss.

The thermogravimetric analysis was performed to evaluate the thermal stability and decomposition behavior of the CP6 base hydrogel and its indocyanine green and doxorubicin-loaded formulations. All of the samples exhibited a multi-step thermal degradation pattern, visually showing the composition of the PVA-chitosan crosslinked hydrogel network and the presence of two drug molecules.

The CP6 hydrogel showed an initial weight loss below 150 °C, which corresponds to the evaporation of physically bound water. The major weight loss occurred between 200-350 °C, representing the thermal composition of the polymer’s backbone, specifically regarding the breakdown of PVA and chitosan chains, as well as the cleavage of genipin crosslinks.

The drug incorporation was shown to have altered the degradation profile. CP6-ICG displayed a slightly earlier onset of degradation, potentially due to the light sensitivity of the ICG and its destabilizing effect on the hydrogel’s structure at elevated temperatures. Similarly, the CP6-DOX showed an increased weight loss in to 20-250 °C range, which suggests that the interactions between the doxorubicin in the hydrogel may have influenced the crosslink density or stability.

The CP6-DOX-ICG formulation showed the greatest mass loss in the higher temperature range above 350 °C. The lower residual weight observed in CP6-DOX-ICG that’s relative to CP6, supports the hypothesis that drug incorporation decreases the thermal stability of the hydrogel, possibly due to the disruption of hydrogen bonds.

Differential Scanning Calorimetry

Figure 8: The DSC thermograms of the CP6 hydrogel formulations. The CP6 base hydrogel (topleft), CP6-ICG (top right), CP6-DOX (bottom left), and CP6-DOX-ICG (bottom right). All samples display endothermic transitions associated with water evaporation and polymer relaxation. Drug loading does alter the thermal transitions, which suggests changes in crystallinity and crosslinking within the hydrogel.

The DSC was used to show the thermal transitions and the energetic behavior of the CP6 hydrogel and its loaded drug formulations. The DSC curves of all of the formulations display multiple endothermic events and indicate the different thermal transitions.

The first large endothermic peak below about 150 °C observed in all of the formulations does correspond to the evaporation of bound and unbounded water molecules, which is to be expected of the hydrophilic nature of PVA-chitosan hydrogels. Additionally, the thermal transitions between 150 C-300 °C reflect the disruption of crystalline regions. In the CP6 hydrogel, this is likely associated with the breakdown of hydrogen bonding and the crystalline area in PVA and chitosan, as well as the genipin crosslinking sites.

Drug incorporation did influence the transitions. The CP6-DOX and the CP6-ICG show shifts in the second endothermic peak, which suggests that the doxorubicin and indocyanine green interfered with the crystallinity or the crosslinking density. The CP6-DOX-ICG formulations show the most evident intensity change over time, showing reduced crystallinity and increased amorphous character due to the entrapment of the drugs.

These results support the hypothesis that the incorporation of indocyanine green and doxorubicin disrupts the physical formation of the hydrogel, which can influence swelling behaviors, degradation rates, and drug release kinetics. The decreased sharpness and intensity of thermal transitions in the CP6-DOX-ICG formulation indicate an enhanced molecular dispersion of the drugs.

Scanning Electron Microscopy

Figure 9: SEM images of the CP6 hydrogel. The micrographs show the top surface (left) and cross-section (right) morphology at different magnifications. The top view reveals a fibrous and interconnected surface, while the cross-sectional images display a porous internal structure with well-defined pores.

The SEM analysis of the CP6 hydrogel revealed a porous network at both the surface and within the cross-section, which is a characteristic of the freeze-thaw method. The top-view viewimages display the interconnected surface with micro- and nano-scale features, while the cross-section views show the well-distributed pores ranging from approximately 100-300 μm. These pores were supported by thin polymer walls, which indicates a stable structure but a highly permeable scaffold. The observed porosity and interconnected structure are ideal for facilitating drug diffusion and nutrient transport. This confirms the hydrogel’s suitability for localized drug delivery.

Figure 10: SEM image of the CP6-DOX hydrogel. The micrographs show the surface (left column)and cross-section (right column) views at magnifications x50, x150, and x250. The images reveal a porous and irregular structure with increased pore density and variability compared to the base CP6 hydrogel. Pores range from 100 to 300 micrometers and appear more disorganized, which suggests that doxorubicin incorporation alters the internal structure.

SEM imaging of the CP6-DOX hydrogel reveals a porous and heterogeneous microstructure. The top-view images show a higher density of rounded and irregular pores, suggesting that doxorubicin incorporation disrupted the uniformity of the hydrogel structure and potentially interfered with the crosslinking and an increase in free volume. Cross-section images revealed more interconnected pores with thinner walls and some collapsed areas, as noted when the magnification increased. This indicates a softer internal structure. The pore sizes ranged from 100 to 300 micrometers, with increased frequency and irregularity. This may enhance drug diffusion but could also impact mechanical stability. These structural modifications are consistent with the enhancement of drug loading capacity and altered swelling or degradation behaviors that are associated with doxorubicin integration.

Figure 11: SEM image of the CP6-ICG hydrogel. The surface (left) and cross-section (right)images at x50, x150, and x250 magnifications illustrate a more compact and less porous structure compared to the CP6 and CP6-DOX. The Cp6_ICG shows reduced pore size and density, showing structural tightening.

The SEM imaging of the CP6-ICG hydrogel revealed a denser and less porous morphology compared to the CP6 and CP6-DOX formulations. The top-view images showed a compact surface with limited pore openings, which shows that indocyanine green may have disrupted the formation of larger pores or caused partial pore collapse during the freeze-thaw cycles. The cross-section images at x150 and x250 magnifications do show the presence of layers and elongated pores. This could suggest that there is a tighter internal structure that could slow down water uptake and drug diffusion. The reduced porosity and thicker pore walls do show a stabilizing effect of ICG on the polymer, which could enhance structural integrity and potentially limit diffusivity.

Figure 12: SEM image of the CP6-DOX-ICG hydrogel. Surface (left) and cross-sections (right)views at x50, x150, and x250 magnifications show a rough and heterogeneous morphology. The dual-drug formulation exhibits large and irregular pores, indicating disruption of the internal structure and suggesting a disruption of the structure from the incorporation of the ICG and doxorubicin.

The SEM images of the CP6-DOX-ICG hydrogel revealed a heterogeneous and highly porous structure as a result of the combined incorporation of the doxorubicin and ICG. The top-view micrographs show a rough surface with visible grooves and variable pore openings, which indicates the disrupted polymer packing. Cross-section images displayed a well-developed internal structure with large and irregularly shaped pores ranging from 100-300 micrometers in size with thin polymer walls. This structure suggests enhanced porosity and diffusion paths that could facilitate controlled delivery of both agents. However, the uneven pore distribution could lead to a possible compromise in the structure due to the competing interactions of the indocyanine green and the doxorubicin.

Swell Test

Ultra Pure Water

Initial Wt, (g) 5m(g) 300 sec 10m(g) 600 sec 30m(30g) 1800 sec 1hr (g) 3600 sec 2hr(g) 7200 sec 4hr(g) 14400 sec 6hr(g) 8hr(g) 24hr(g) 30hr(g)  48hr(g) 72hr(g)  96hr(g)  120hr(g)  144hrs 7days 8days 9days 10days
CP6 0.0197 0.1168 0.1582 0.2000 0.2145 0.2508 0.2689 0.2975 0.3188 0.3647 0.3797  0.4095  0.5245  0.5500  0.5735  0.6086 0.6357 0.6423 0.6077 0.5677
CP6 ICG 0.0202 0.1611 0.2007 0.2465 0.2632 0.2972 0.3042 0.3215 0.3487 0.4208 0.4316  0.4651  0.5059  0.5297  0.4850  0.4644 0.5181 0.5179 0.5175 0.5056
CP6 DOX 0.0184 0.1309 0.1634 0.2055 0.2182 0.2366 0.2523 0.2753 0.2825 0.3455 0.3528  0.3842  0.4439  0.4626  0.4816  0.4798 0.4845 0.4700 0.4782 0.4642

Figure 13: Swelling behavior of the CP6 hydrogel formulations in ultra-pure water over 10 days. Initial dry weights were recorded, and hydrogels were submerged in ultra-pure water at multiple time intervals from 5 minutes to 10 days. The base CP6 hydrogel exhibits the highest overall swelling capacity.

Swelling analysis in ultra-pure water shows the absorption behaviors among three different formulations. CP6 displays the greatest swelling capacity, increasing from an initial weight of 0.0197 g to a peak of 0.6423 g by day 8, indicating a high water affinity and extensive pore volume.CP6-ICG also showed significant swelling, but it plateaued earlier and reached a slightly lower maximum (0.581g at day 7), suggesting that ICG loading may have partially reduced free hydrophilic sites. CP6-DOX showed the lowest swelling capacity with a maximum of 0.4845 g, potentially due to an increased crosslink density. All formulations showed a rapid initial uptake within the first few hours, followed by a slower equilibration phase, consistent with typical swelling kinetics in porous hydrogels.

Pure DMEM

Initial Wt, (g) 5m(g) 300 sec 10m(g) 600 sec 30m(30g) 1800 sec 1hr (g) 3600 sec 2hr(g) 7200 sec 4hr(g) 14400 sec 6hr(g) 8hr(g) 24hr(g) 30hr(g)  48hr(g) 72hr(g)  96hr(g)  120hr(g)  144hrs 7days 8days 9days 10days
CP6 0.0197 0.1168 0.1582 0.2000 0.2145 0.2508 0.2689 0.2975 0.3188 0.3647 0.3797 0.4095 0.5245 0.5500 0.5735 0.6086 0.6357 0.6423 0.6077 0.5677
CP6 ICG 0.0202 0.1611 0.2007 0.2465 0.2632 0.2972 0.3042 0.3215 0.3487 0.4208 0.4316 0.4651 0.5059 0.5297 0.4850 0.4644 0.5181 0.5179 0.5171 0.5056
CP6 DOX 0.0184 0.1309 0.1634 0.2055 0.2182 0.2366 0.2523 0.2753 0.2825 0.3455 0.3528 0.3842 0.4439 0.4626 0.4816 0.4798 0.4845 0.4700 0.4782 0.4642

Figure 14: Swelling behavior of CP6 hydrogel formulations in pure DMEM over 10 days. CP6, CP6-ICG, and CP6-DOX hydrogels were submerged in Dulbecco’s Modified Eagle Medium(DMEM), and swelling was monitored over 10 days. Unlike in ultra-pure water, all formulations exhibited limited and declining swelling trends after initial uptake, with CP6 showing slightly higher absorption than the drug-loaded variations. The presence of salts and proteins in DMEM may have influenced osmotic balance.

In contrast to the swelling observed in ultra-pure water, all hydrogel formulations showed reduced swelling in pure DMEM, with peak values within the first hour and subsequent weight loss over time. CP6 absorbed the most overall, increasing from 0.0234 g to an early maximum of 0.1276 g at 10 minutes before gradually declining. CP6-DOX followed a similar trend, reaching a peak of 0.1063 g, while CP6-ICG demonstrated the lowest and most rapid deswelling since it never exceeded 0.0876 g. This behavior is due to the ionic crosslinking and osmotic imbalances due to the presence of salts, amino acids, and glucose in DMEM. These findings suggest the physiological environment significantly alters hydrogel swelling.

PBS 7.4

Initial Wt, (g) 5m(g) 300 sec 10m(g) 600 sec 30m(30g) 1800 sec 1hr (g) 3600 sec 2hr(g) 7200 sec 4hr(g) 14400 sec 6hr(g) 8hr(g) 24hr(g) 30hr(g)  48hr(g) 72hr(g)  96hr(g)  120hr(g)  144hrs 7days 8days 9days 10days
CP6 0.0218 0.0857 0.1032 0.1338 0.1441 0.1591 0.1668 0.1722 0.1806 0.1994 0.1938 0.2102 0.1963 0.2047 0.1980 0.1254 0.1336 0.1343 0.1332 0.1360
CP6 ICG 0.0170 0.0915 0.1131 0.1355 0.1489 0.1600 0.1617 0.1685 0.1820 0.1993 0.2038 0.2116 0.1235 0.1271 0.1296 0.1359 0.1300 0.1297 0.1444 0.1394
CP6 DOX 0.0153 0.0355 0.0523 0.0378 0.0478 0.0381 0.0371 0.0403 0.3520 0.0386 0.0449 0.0347 0.0378 0.0383 0.0418 0.0427 0.0413 0.0394 0.0420 0.0402

Figure 15: Hydrogels CP6, CP6-ICG, and CP6-DOX were immersed in PBS and monitored at regular intervals for 10 days. CP6 and CP6-ICG exhibited gradual and stable swelling, reaching maximum uptake between 24-72 hours. CP6-DOX showed significantly reduced swelling behaviors, indicating altered structural response in a buffered ionic environment.

Swelling analysis in PBS (pH 7.4) revealed that both CP6 and the CP6-ICG hydrogels maintained moderate and consistent water uptake, peaking around 72 hours (0.2102 g and 0.2116 g) before stabilizing with minor fluctuations. This suggests that the ionic strength of PBS did not strongly inhibit swelling for these formulations. The CP6-DOX shows reduced swelling throughout the entire duration, with irregular absorption and values remaining below 0.045 g, except for a single outlier at the 8-hour mark. The suppressed swelling could be due to interactions between the doxorubicin and the chitosan backbone, additionally leading to reduced porosity. These findings show that buffer composition and drug loading significantly affect hydrogel swelling dynamics.

 

Alamar Blue

Figure 16: Bar graph displays the percentage of viable cells following the treatment with unloaded and loaded hydrogels: CP6, CP6-ICG, CP6-DOX, CP6-DOX-ICG, with or without near-infrared irradiation exposure, as well as controls and insert-based delivery systems. Untreated control samples show consistent high viability, while free DOX and DOX-ICG formulations significantly reduced cell viability, especially at 48 and 72 hours. CP6 alone exhibited minimal cytotoxicity, while insert-based and NIR assisted formulations demonstrated time-dependent effects on metabolic activity, which indicates enhanced photothermal or chemotherapeutic impact in the CP6-DOX-ICG.

Alamar Blue assay reveals distinct cytotoxicity properties amongst the various hydrogel formulations and treatment conditions over time. The CP6 hydrogel showed minimal cytotoxicity, maintained cell viability above 30% at 24 hours, and returned to baseline at later time points. This confirms its biocompatibility. In contrast, the formulations containing doxorubicin showed pronounced cytotoxic effects, with viability dropping below 1% by 48 hours, which highlights doxorubicin’s chemotherapeutic activity. Interestingly, ICG alone had relatively mild effects on cell viability, while the NIR exposure did slightly enhance the cytotoxicity in the ICG and DOX-formulated samples, likely due to the photothermal effects. The cell culture insert samples showed a more gradual and sustained viability reduction, specifically with the ICG and DOX formulations, which suggests prolonged release and localized accumulation. The most significant reduction in inviability occurred in DOX-ICG NIR insert groups. These findings demonstrate that the CP6-DOX-ICG hydrogels, especially under NIR activation, are highly effective in reducing cell viability, while CP6 alone remains non-toxic and suitable to be a carrier for controlled drug delivery.

Conclusion

This study demonstrates the successful development and characterization of genipin-crosslinked PVA-chitosan hydrogels for the potential application in localized melanoma therapy. Structural analysis via SEM confirmed the presence of pores, with the drug incorporation controlling the distribution. TGA and DSC revealed that drug loading altered the thermal stability and phase behavior of the hydrogels, while the tensile testing confirmed that the mechanical integrity was retained after crosslinking. Swelling studies conducted in ultra-pure water, PBS, and DMEM highlighted that CP6-DOX-ICG and CP6-DOX have the lowest swelling capacity, likely due to the drug and polymer interactions. Importantly, the Alamar Blue assays confirmed the biocompatibility of CP6 alone, while DOX and ICG-loaded formulations, especially the samples exposed to NIR, showed significant cytotoxicity toward melanoma cells, which supports therapeutic potential. While the results of this study are promising, the research does remain ongoing, and further in vitro assays are currently being on to assess biocompatibility and drug diffusion.

References

1. Singh, P., Raj, K., Shrivastav, A., & Sharma, V. (2024). Innovative biodegradable polymer hydrogel beads for enhanced controlled drug delivery: Formulation and characterization. Biosciences Biotechnology Research Asia, 21(4), 1451-1462 https://www.biotech-asia.org/vol21no4/innovative-biodegradable-polymer-hydrogel-beads-for-enhanced-controlled-drug-delivery-formulation-and-characterization/

2. Yu, S., He, C., & Chen, X. (2018). Injectable hydrogels as unique platforms for local chemotherapeutics-based combination antitumor therapy. Macromolecular Bioscience, 18,1800240. https://onlinelibrary.wiley.com/doi/10.1002/mabi.201800240

3. Elhabal et al. (2024). Enhancing Photothermal Therapy for Antibiofilm Wound Healing. International Journal of Nanomedicine.

4. Ahmed et al. (2023). Role of Thermal and Reactive Oxygen Species-Responsive Synthetic Hydrogels in Localized Cancer Treatment. Materials Advances.

5. Zheng et al. (2015). Ultra-Small Mesoporous Silica Nanoparticles as Efficient Carriers for pH-Responsive Releases of Anti-Cancer Drugs. Dalton Transactions.

6. Skin Cancer Foundation. (2023). Skin cancer facts and statistics. Retrieved from https://www.skincancer.org/skin-cancer-information/skin-cancer-facts/

Research in Chemistry and Biochemistry

Synthesis and Characterization of Gold Nanorods and Gold Nanorod Dimers  

By Kayli Vieira 
Introduction 

Gold nanoparticles have become a hot spot in recent research due to their unique electrochemical properties. When light hits the surface of a gold nanoparticle, a collective oscillation of electrons results in a strong electromagnetic field, enhancing absorption and scattering properties. This phenomenon, known as plasmon resonance, gives gold nanoparticles their uniqueness and versatility in different areas of research. For example, gold nanoparticles have been utilized for research in drug delivery, biosensing, spectroscopy, and catalysis. In Professor Wei-Shun Chang’s lab at UMass Dartmouth, gold nanoparticles are studied using spectroscopic techniques to understand their physical and chemical properties for application in catalysis, microscopy, and biosensing. For this project, gold nanorods will be used to study substrateinduced chirality through circular dichroism (CD) measurements on a hyperspectral microscope. Samples of both chiral and achiral gold nanorods are required for single particle measurement, which can be synthesized using simple wet-lab procedures.  

Results 

The first step of the project requires the synthesis of pure, homogenous, mediumsized gold nanorods (22×66 nm in size). To accomplish this, multiple syntheses were performed using the seed-mediated growth method first developed by the Murphy Group (see Figure 1 for procedure schematic). 

Figure 1. Schematic of Seed-Mediated Growth Method of Gold Nanorods (Adapted from the Murphy Group) 

After extensive background research and alterations of reagent concentrations, the procedure was used to synthesize mediumsized gold nanorods, optimized for single particle measurement. For characterization, samples were analyzed using UV-Vis spectroscopy, scanning electron microscopy, and hyperspectral microscopy.  

Figure 2. UV-Vis Absorption Graph of Medium Sized Gold Nanorods

As seen in Figure 2, the UV-Vis absorption graph yielded important information about the sample’s purity and general aspect ratio, evident by the transversal and longitudinal plasmon bands. To determine the average size of the nanorods produced, SEM (scanning electron microscope) images were taken and analyzed. After analyzing over 150 individual gold nanorods, the dimensions of the rods produced yielded 59.6 7.0 nm X 20 3.18 nm (image can be seen in Figure 3).  

Furthermore, after being spin cast on a glass substrate and imaged on a darkfield inverted hyperspectral microscope, individual spectra for a given nanorod were produced using analyses of Hyperspectral images through the MATLAB program (see Figures 4 and 5). 

Figures 3 and 4. SEM Image and Hyperspectral Image of Medium Gold Nanorod Dimers 

Figure 5. Individual Gold Nanorod Spectra 

After successful synthesis of singular mediumsized gold nanorods, synthesis of a chiral sample is required. Using the previously made gold nanorod samples and a method developed by Kar et. al, samples of gold nanorod dimers (two gold nanorods linked end-to-end) were assembled. Using the polyelectrolyte coating and thiol-linking procedure schematized in Figure 6, chiral samples of gold nanorod dimers were synthesized and stable for a very short time.

Figure 6. Schematic of Gold Nanorod Dimers developed by Kar et. al 

Unfortunately, results varied from those in the literature. After a successful first attempt, reproducing the same results was almost impossible. Countless experiments were done altering individual factors at every part of the procedure to determine the driving force of the reaction. The hydrochloric acid concentration was determined to be the main contributor as to whether dimers were produced, likely due to the decrease of surface charge on the rods resulting in the ability for linkage. Furthermore, it was discovered that the addition of DABCO resulted in the destabilization of dimers formation likely due to surface charge imbalances. Thus, successful samples produced were only stable for a few hours.  

Figure 7. UV-Vis Absorption Graph of Gold Nanorod Dimers Sample

Characterization of each gold nanorod dimer sample was preformed using the same methods mentioned above. In Figure 7, the UV-Vis absorption graph revealed important information as to whether or not dimers were formed in the reaction, evident by the shoulder created after the longitudinal plasmon band. Similar graphs can be seen in the resulting spectra generated using the hyperspectral microscope (see Figures 8 and 9). Furthermore, SEM images depicted in Figure 10 clearly showed the configuration of the rods as dimers, aligning with the results of the hyperspectral images and UV-Vis graph. 

Figures 8 and 9. SEM Image and Hyperspectral Image of Medium Gold Nanorod Dimers

Discussion and Future Work 

Homogeneous and pure medium-sized gold nanorods were successfully synthesized for future single-particle measurement. Additionally, gold nanorod dimers were assembled using the polyelectrolyte coating and thiol linking method previously developed in the literature. Hydrochloric acid was identified to be the key parameter of the dimerization reaction, with alterations needed for each concentration of rods used for assembly. Currently, the gold nanorod dimers synthesis is continuing to be optimized for maximum reproducibility, with future plans of single-particle measurement. Both products synthesized will be used for analysis in the future study of substrate-induced chirality.  

Works Cited

Synthesis of Solution-Stable End-to-End Linked Gold Nanorod Dimers via pH-Dependent Surface Reconfiguration Ashish Kar, Varsha Thambi, Diptiranjan Paital, Gayatri Joshi, and Saumyakanti Khatua Langmuir 2020 36 (33), 9894-9899 DOI: 10.1021/acs.langmuir.0c01516

Seeded High Yield Synthesis of Short Au Nanorods in Aqueous Solution Tapan K. Sau, and Catherine J. Murphy Langmuir 2004 20 (15), 6414-6420 DOI: 10.1021/la049463z

Research in Biology

Phosphorus Removal Through Plant Assimilation in Floating Treatment Wetlands

By Mia Oliveira, Sara Sampieri Horvet, Micheline Labrie 

Affiliation: Coastal Systems Program, Department of Estuarine and Ocean Science, School for Marine Science and Technology 

Abstract 

Excess phosphorus (P) in freshwater systems contributes to the degradation of aquatic habitats through algal blooms and low dissolved oxygen levels. This study investigates the deployment and effectiveness of floating treatment wetlands (FTWs) as a method for phosphorus removal in Long Pond, located in Barnstable, MA. Preliminary results demonstrate species-specific differences in P assimilation and overall plant biomass increase during the first year of growth. 

Introduction 

Freshwater ponds and lakes in Massachusetts are degraded by excess inputs of phosphorus (P). Bioavailable dissolved forms of P, such as ortho-phosphate (PO43-), can stimulate phytoplankton blooms, potentially containing toxins. These blooms lead to low water column dissolved oxygen, resulting in poor habitat health. 

Long Pond (Barnstable, MA) is currently impaired by excessive P loading, primarily from wastewater sources (Eichner et al. 2022). Although sewering is expected to significantly reduce P inputs, it will not be implemented for ~25 years. The Town of Barnstable is exploring low-cost, non-infrastructural alternatives that can be implemented within 5 years. One such option is the deployment of floating treatment wetlands (FWTs). FTWs transform bioavailable P into plant biomass, which can be permanently removed upon harvest, effectively reducing P levels in the water (Lane et al. 2016). 

Using guidance from the Cape Cod Commission (Eichner et al. 2003), the target total P in the water column is 7.4 kg, consistent with levels measured in 2013 (7.8 kg) and 2011 (6.4 kg). This target contrasts sharply with the 2021 measurement of 16.2 kg, reflecting the need for additional water quality data collection and P reduction. 

Objective: Quantify in-pond P removal by FTW plant biomass via assimilation by seven aquatic plant species. Ultimately, success will be evaluated based on total P removal (kg/year), in situ P reduction from water quality measurements, and cost per kg P removal for full-scale implementation. 

Figure 1.

(Left) Aerial image of Long Pond, a relatively shallow, ~50-acre Great Pond. It is a public resource and subject to MA and federal regulations. The gray square represents the FTWs deployment location. Blue stars represent 2024 water quality sampling sites.

(Right) September 2024 Photo of the FTW rafts. 

Materials and Methods
Study Site

Ten FTW rafts were installed in the northern section of Long Pond in April 2024. Each raft measured 2.0 m × 2.1 m and consisted of a buoyant, high-density polyethylene matrix with designated planting holes for 36 (n=6) or 72 (n=4) plants per raft (Figure 2). The selection of plant species and planting configurations was directed by the Town of Barnstable, with 5 cm plugs sourced from New England Wetland Plants, Inc. 

Figure 2. Aquatic plant species and planting configuration determined by the town. The town obtained 5 cm plugs from New England Wetland Plants, Inc. 

Plant biomass measurements and tissue samples were collected in May (pre-planting) and September 2024 (end of Year 1 growing season). Sub-samples included whole plant wet weight, and wet and dry weights of above and below-water biomass. At the Year 1 season’s end, the ten rafts were surveyed to determine plant survival. 

Biomass was dried to a constant weight of 65°C and ground using a ball mill. Total P was measured via persulfate digestion and analyzed by the ascorbic acid-molybdenum blue method (Method 4500-PE) using a spectrophotometer set at 882 nm at the Coastal Systems Program Analytical Facility. 

Figure 3. Mia Oliveira conducting persulfate digestions at the Coastal Systems Program, SMAST-West. 

Results 

Plants deployed in rafts with 36 planting holes showed slightly higher survivability compared to rafts with 72 planting holes (Table 1). Preliminary data from the town indicated that most plant mortality occurred during July. Higher plant mortality coincided with greater incidence of weed growth. Species with higher survival rates contributed proportionally more to total P removal. 

Table 1. Summary of plant species remaining at the end of Year 1 for each raft. On average, rafts had 67% and 55% survival with 36 and 72 planting holes, respectively. 

 

Species 

Raft # 
36 planting holes  72 planting holes 
1  2  3  4  5  6  7  8  9  10 
C. lurida  9  5  9  2  8  17  9  0  7  14 
J. effusus  6  8  8  10  8  3  9  21  13  11 
P. cordata  0  0  0  0  0  0  0  0  0  0 
A. americanus  3  2  2  4  13  3  5  10  11  10 
A. incarnata  4  5  5  2  0  0  12  8  1  8 
S. atrovirens  3  0  0  0  0  0  1  0  0  0 
V. noveboracensis  0  0  1  2  1  1  2  0  3  4 
Percent Survival  69%  56%  69%  56%  83%  67%  53%  54%  49%  65% 

Initial P content varied across plant species (Figure 4). Acorus americanus displayed the highest P concentration in shoot biomass and significant root P content. Viburnum noveboracensis exhibited higher root P levels compared to shoots, though data was limited (n=1). Comparing initial and final phosphorus contents, phosphorus (umol/g dry plant biomass) decreased over the growing season. On average, root and shoot phosphorus content decreased by 58% and 52%, respectively. 

Whole plant biomass increased across all species from deployment to the end of Year 1, except for P. cordata, which showed no growth. C. lurida had the highest biomass increase, reaching 500 g wet weight. Sub-sampling in September 2024 identified a total of 80 C. lurida individuals, accounting for a 37 kg biomass increase. Similarly, J. effusus contributed 30 kg of biomass over the growing season. Overall, given the wet biomass increase (Figure 5) and the total number of plants remaining at the end of the growing seasons, the total wet biomass increase was approximately 100 kg. 

Figure 4. Total phosphorus content (umol/g; dry wt. basis) of plant roots and shoots subsampled at the time of deployment (a. Initial) or at the end of the Year 1 growing season (b. Final). 

Figure 5. Whole plant biomass (g wet wt.) measured at the time of deployment (May 2024) and at the end of the first year growing season (September 2024). 

Discussion 

The results highlight the different phosphorus uptake capacities among studied aquatic plant species. The growth and P assimilation data demonstrate the potential effectiveness of FTWs as a P mitigation strategy. However, the feasibility of FTWs will likely depend on the biomass increases observed during the second year growing season.

C. lurida and J. effusus showed promise for selection for future FTW implementations, given their biomass increase. It is likely that biomass increase, rather than phosphorus content, will control total P assimilation. The lower performance of P. cordata suggests that careful species selection is important for optimizing FTW efficiency. Factors such as planting density, environmental stressors (competition, predation), and species-specific nutrient uptake mechanisms likely influenced survival and growth outcomes.

Additionally, the cost-effectiveness of FTWs should be assessed relative to other mitigation strategies. Integrating FTWs with other methods, such as aluminum sulfate treatment, sediment dredging, or aeration systems, could provide a more comprehensive and reliable approach to P management in Long Pond. Future research should focus on long-term monitoring of FTWs, including multi-year survival rates, biomass increase potential, and the ecological impacts of large-scale deployments. Additional work is also needed to quantify in situ water column P reductions and evaluate potential secondary benefits, such as enhanced habitat for aquatic organisms and sedimentation. 

Acknowledgements 

We thank the Office of Undergraduate Research and the Town of Barnstable Department of Public Works for their invaluable support. Special thanks to the individuals and organizations dedicated to the protection and restoration of Barnstable’s aquatic ecosystems.  

References

Eichner, E.M., T.C. Cambareri, G. Belfit, D. McCaffery, S. Michaud, and B. Smith. (2003).

Cape Cod Pond and Lake Atlas. Cape Cod Commission. Barnstable, MA.

Eichner, E., Howes, B., & Schlezinger, D. (2022). Long Pond management plan and diagnostic assessment. Town of Barnstable, Massachusetts. Coastal Systems Program, School for Marine Science and Technology, University of Massachusetts Dartmouth. New Bedford, MA. 110 pp.

Lane, S., Sample, D., Lazur, A., Winston, R., Streb, C., Ferrier, D., Linker, L., & Brittingham, K. (2016). Recommendations of the expert panel to define removal rates for floating treatment wetlands in existing wet ponds. Final Report to the Urban Stormwater Work Group.

Standard Methods for the Examination of Water and Wastewater, 19th edition. Method 4500- PE.

Research in Mechanical Engineering

Mixed Mode Fracture Characterization of Hydrogel-3D Printed Polymer Adhesively Bonded Systems 

By Nicholas Sardinha

Introduction 

Hydrogels are soft polymer networks with an interstitial liquid, usually water, and they have many applications ranging from tissue engineering, water purification [1], and even anti- biofouling. However, for hydrogels to be more useful for anti-biofouling, their adhesion to different surfaces must be investigated. Many marine vessels have hulls made of different metals such as steel and aluminum, but with the rise of additive manufacturing came the increase in the use of 3D-printed polymers as hull materials. This research attempted to study the adhesive bonding of hydrogel and 3D-printed acrylonitrile butadiene styrene (ABS). The specific method chosen to evaluate the fracture toughness was the J-integral method, used on a mixed mode bending fixture described in the methodology section of this report and based on the work of Zhao Y, Seah LK, Chai GB [2]. Mixed mode bending applies opening mode-I and shearing mode-II to the double cantilever beam (DCB) specimen. However, this type of test has never been conducted with two materials with such a large difference in elastic moduli, with hydrogel having a Young’s modulus in the 10 kPa – 2 MPa range and ABS in the 2-4 GPa range. This causes the ABS to experience much less deformation than the hydrogel at the same load. That is the primary reason why the traditional peel test was not used, as hydrogel experiences significant deformation, which absorbs energy. This leads to the peel test not accurately measuring fracture toughness of softer materials, since that additional energy absorption is not properly accounted for in fracture toughness calculations. 

Methodology 

Specimen Production 

Table 1: Hydrogel Composition 

Number of Samples  Total Solution (ml)  DI Water (ml)  AAm (g)  DAC (ml)  MBAA 

(mg) 

KA (mg) 
1  25  14.8  8.9  10.2  23.2  13.7 
3  75  44.4  26.7  30.6  69.6  41.1 

The specimens used in this research were fabricated using a unique in situ fabrication method. Firstly, a hydrogel precursor solution was made using the five main components: deionized water (DI Water), Acrylamide (AAm) monomer, Acryloyloxyethyltrimethyl ammonium chloride (DAC) cationic monomer, N,N’-Methylenebisacrylamide (MBAA) crosslinker, and α-Ketoglutaric acid (KA) photoinitiator. From Table 1 the weight percentage of AAm and DAC are 25.24 wt.% and 32.69 wt.%, respectively. This process produces a clear hydrogel precursor solution that can be used for the fabrication of test specimens. 

The second half of each specimen is a 3D-printed ABS T-shape, which was produced using a Stratasys uPrint 3D printer and ABS P430. After the ABS T-shape finished printing, a 10 mm long pre-crack film was glued to the top surface, as shown in the red area of Figure 1(a). 

 For the control samples, the rest of the 3D-printed surface was left unaltered, and the ABS T- shape was placed into a silicone mold. This silicone mold was designed with two connected cavities: one for the ABS and the other for the hydrogel precursor solution. The side for the precursor solution had an opening that was sealed using a strip of photocopier film and two glass microscope slides. The purpose of using transparent materials to seal that side of the mold is to ensure that the 365 nm ultraviolet light can pass through and photocure the hydrogel. After many experiments, a 4-hour photocuring time was chosen and used for all samples. Since the mold had an opening on the top for the hydrogel precursor solution to be poured in, a sheet of plastic wrap was applied around the entire mold to prevent moisture from entering or exiting during curing. Photocopier film was also used to line the inside of the mold to ensure that the surface of the hydrogel would be smoother. 

After the samples finished curing, they were removed from the mold, and excess hydrogel was cut away with a razor blade so that the edge of the hydrogel aligned with the edge of the ABS. Once this was completed, a 20× 100 mm2 strip of photocopier film was superglued to the backside of the hydrogel. Next, a 100 mm long strip of paper with millimeter markings was applied to the side of the ABS to help measure the pre-crack length. To ensure that the pre- crack area did not have any adhesion, the crack was slightly opened by hand, and the end of the pre-crack was marked on the side of the ABS with a black marker. 

Photooxidation 

The photooxidation process refers to the exposure of the ABS surface to chlorine dioxide (ClO2) gas and 365 nm ultraviolet light. To produce ClO2, 100 mg of sodium chlorite (NaClO2) was added to 14 ml of DI water inside a small dish. This small dish was then transferred into a larger glass reaction vessel. Any ABS samples to be photooxidized were then placed in the reaction vessel close to the small dish. The entire reaction vessel was then placed in the fume hood. Then using an adjustable pipette, 100 µl of 37% hydrochloric acid (HCl) was added to the solution, which starts producing chlorine dioxide gas. After that, the reaction vessel was sealed using plastic wrap and a clear flat plastic sheet to contain all the produced gas. Then the ultraviolet light was placed on top of the reaction vessel, and a timer was started as soon as the light was switched on. After the desired time was completed, the ABS samples were transferred into another container, where they could be used immediately for either contact angle measurements or for fabrication of mixed mode bending specimens. All remaining reacted solution was then transferred into a designated waste container specifically for that solution. 

Mixed Mode Bending Fixture 

Figure 1:(a) Hydrogel/ABS adhesive system and (b) T-Shape drawing with dimensions in millimeters. 

 

Figure 1(a) shows the ABS/hydrogel adhesive systems, and Figure 1(b) shows the dimensions of the T-shape part of either portion. The double T-shape specimens were placed into a mixed-mode bending fixture as shown below in Figure 2. This fixture was loaded under compression with a load P, where force and displacement measurements were recorded using a Shimadzu AGS-X universal testing machine (UTM) with a 50N load cell. A Celestron handheld digital microscope was used to capture the crack initiation. The loads at locations A, B, C, and D were determined using statics, and the angles measured at these four locations were used in determining the fracture toughness. Due to the significant difference in elastic moduli between the hydrogel and ABS, θB & θD are extremely small and approximated as zero. 

 

Figure 2: Mixed mode bending loading configuration of the adhesive system. 

 

The lever length c can be adjusted to change the mode mixity, and for these experiments a length of 30.61 mm was used. The horizontal distance between points B and C called length L in Figure 2 cannot be adjusted and was set to 45 mm. Prior to each set of tests, the load 50 N load cell was calibrated. 

To secure the prepped samples into the fixture, superglue was applied to the top surface of the hydrogel to meet the grip, and the grips were tightened down. The lower grips on the ABS side did not require superglue, since the ABS did not deform during testing. Once the specimen was in the fixture, the pivot was slowly lowered until it contacted the surface, then a vertical line was drawn onto the hydrogel to measure the angle at point C. Both the force and displacement were then zeroed. A loading rate of 4 mm per minute was used based on previous mixed mode tests. The recording on the digital microscope and the UTM were started at the same time and two screenshots were taken, one at the start of the experiment and one at fracture initiation. 

These screenshots were then imported into MATLAB, where a code was used to analyze the angles of 3 different points. The first screenshot was analyzed to measure the initial angle of the vertical line at point C, and the second screenshot was analyzed to measure the angle of the fixture, the angle at A, and the angle at C (this is shown below in Figure 3). The time of fracture initiation was also recorded from this video and was used to get the load and displacement at fracture initiation. With the P, θA, & θC, the fracture toughness can be calculated using the following formula:

= 100( sin() + sin() sin() + sin() ) 

The average fracture toughness and 95% confidence interval were plotted using Excel. The force vs displacement was also plotted using MATLAB and the Excel spreadsheets produced by the UTM. After the completion of each test, the hydrogel was separated from the ABS and the ABS surface was cleaned using deionized water for future use. This procedure was repeated until four consistent tests were performed. 

Figure 3: Example of Angle Measurements at Fracture Initiation. 

 

Water Contact Angle Measurement 

Water contact angle measurements were taken using a Ramé-Hart Model 90 goniometer. The goniometer was calibrated using the included calibration sphere and the water contact angle software package. The surface to be tested was placed on the test platform with the backlight turned on and leveled using the included adjustment knobs. After that, a droplet of deionized water was dropped on the surface using the included syringe, and the contact angle was recorded. This was repeated in four separate locations on the ABS surface on two different specimens. This procedure was repeated for the three different conditions which were untreated, 15-minute photooxidation, and 30-minute photooxidation. The average and 95% confidence interval were then plotted in Excel. 

Results 

Figure 4: Water Contact Angle for Untreated, 15 Minute Photooxidation, and 30 Minute Photooxidation. 

 

The water contact angle measurements from Figure 4 showed a significant difference between untreated samples and samples that have had the photooxidation procedure done on them. The average contact angle decreased from 76.23° to 42.06°. This was an expected result as the photooxidation process modifies the ABS surface with more reactive groups, meaning the surface should become more hydrophilic, which it did. It also showed that increasing the photooxidation treatment time from 15 minutes to 30 minutes did not have much of an effect. This is important information as it led to one of the challenges experienced later on in the process. 

 

Figure 5: Control Load vs Displacement. 

 

The load versus displacement graph in Figure 5 shows the same trend for all specimens. There is an initial shallower slope section that quickly transitions into a linear portion until the peak load is achieved, and then the load quickly drops off as the crack grows along the interface. All four specimens achieved peak load values within 1 N of each other. 

Figure 6: Control Fracture Initiation Toughness. 

 

The average fracture toughness for the untreated control samples was 31.1 J/m2 as seen in Figure 6. There are no other conditions to compare to the control which is explained in the discussion section of this report. 

Discussion 

Many challenges were faced during this research that limited the scope of this project. Due to the novelty of bonding hydrogel and ABS plastic, a unique hydrogel precursor solution had to be created and iterated on. Initially, the hydrogel used had half of the concentration of AAm and DAC, which caused the hydrogel to be very soft, excessively deform, and ultimately lead to inconclusive tests. This challenge was eventually overcome by doubling the concentration of AAm and DAC, along with increasing the MBAA cross-linker concentration. 

This led to very successful control experiments. However, after the control tests concluded, photooxidation tests started. Based on the contact angle results, a 10-minute photooxidation procedure was selected. As expected, the photooxidation had a major effect on the adhesive bond between the surfaces; it had such a large effect that the hydrogel consistently failed prior to the interfacial bond. Corrective measures were attempted, such as decreasing the concentration of the NaClO2 and HCL solution by half and the exposure time down to 3 minutes. This still resulted in the same challenges as previous tests, namely crack propagation into the hydrogel. 

The second-largest challenge was determining the proper method to measure θC. The methods used by other researchers who used the same fixture did not work as expected due to the ABS acting as a rigid body and causing the hydrogel to indent rather than bend. This indentation issue was compounded by the fact that the hydrogel used in these experiments was prone to warping. Many different solutions were proposed to combat those problems, but ultimately, applying a backing layer to the hydrogel and drawing a vertical line at point C was chosen, as described in the methodology section of this report. 

Conclusion / Future Work 

There are many more aspects of this research topic that can be further investigated, the most pressing being the degree to which the photooxidation process increases the fracture toughness. This would require the control samples to be redone using a hydrogel stronger than an interfacial bond to prevent cracks from propagating into the hydrogel. Once the hydrogel is strong enough to withstand the photo-oxidized tests, other test configurations can be tested, such as modifying the surface geometry or roughness. Following that, another future step for this research is using digital image correlation (DIC) to determine the mode mixity of these mix mode tests. The lever length used for these tests was chosen because, with the same material for top and bottom T-shapes, it produces equal parts mode I & II loading conditions. However, since the top and bottom are not the same material, the mixity is not the same. With the help of DIC, the strain fields near the crack tip can be analyzed to determine the true mode mixity. Overall, this research has built up some knowledge of hydrogel and ABS adhesively bonded systems. Even though it did not achieve all of its goals from its onset, it still provides a solid foundation that can be built upon by future researchers. 

References 

Yahia LH, Chirani N, Gritsch L, et al. History and Applications of Hydrogels. J Biomedical Sci. 2015, 4:2. http://dx.doi.org/10.4172/2254-609X.100013. 

 Y. Zhao, L. K. Seah, G. B. Chai, Measurement of interlaminar fracture properties of composites using the J-integral method, J. Reinf. Plast. Compos. 35 (2016) 1143-1154. https://doi.org/10.1177/0731684416642031. 

 

Research in English

Evolving Feminism in The Hunger Games Series

By Ellie Cook

My project analyzes the differences in feminine gender performance between Katniss Everdeen, the protagonist of the original The Hunger Games trilogy by Suzanne Collins, and Lucy Gray Baird, the protagonist of the prequel novel The Ballad of Songbirds and Snakes. It analyzes the two characters from an intersectional perspective and argues for the importance of cultural, economic, social, etc. contexts when accounting for differences in gender performance, as these factors are all intertwined with gender norms. From the basis of their contrasting gender performance, the paper then delves into agency as a theme and how each character’s relative agency impacts their feminine representation. These differences in gender performance and agency show a different archetype for a young adult feminine heroine over ten years after the original trilogy’s publication. 

From the end of May to mid-July, I completed the background reading to prepare for writing my paper. I read critical theory such as Judith Butler’s Gender Trouble, Undoing Gender, and Who’s Afraid of Gender? as well as Butler Matters: Judith Butler’s Impact on Feminist and Queer Studies by Warren J. Blumenfeld and Margaret Sönser Breen. I also read scholarship about The Hunger Games series that focused on gender and feminism. Some of these texts include: “You Love Me. Real’: Gender in the Hunger Games Trilogy” by Bienvenue Bray and “I Hunt. He bakes.’: Constructing and Deconstructing Gender Identity in Suzanne Collins’s The Hunger Games Trilogy” by Rakchuda Thibordee. I also read books that gave a general overview of the young adult dystopian genre, such as Female Rebellion in Young Adult Dystopian Fiction by Sarah K. Day, et al. and Contemporary Dystopian Fiction for Young Adults: Brave New Teenagers by Carrie Hintz, et al. Using these texts, I completed an annotated biography.

I met with my Apex mentor, Professor Caroline Gelmi, in mid-June, mid-July, and mid-August to check my progress. From mid-July to the end of August, I completed the first draft of my paper. I will continue to work on it with Professor Gelmi throughout the Fall 2024 semester and will finish my final draft at the end of the semester.  

I am grateful to Professor Gelmi for her feedback and mentorship and to OUR for supporting this project through a summer research grant.

Research in Chemistry and Biochemistry

Microwave Mediated Synthesis of Quinazolinone Natural products of  Peganum harmala for Medicinal Chemistry Applications 

By Amelie Duval
Introduction

In the pharmaceutical field, quinazolinones have become an important pharmacophoric scaffold due to their presence in natural compounds with a wide range of medicinal chemistry applications. Among these, Peharmaline A (1), a natural alkaloid with b-Caroline and tricyclic pyrroloquinazolinone cores (Figure 1) that exhibits various biological activities, including antimalarial, anticancer, anti-inflammatory, and antibacterial activities has garnered significant interest of synthetic and medicinal chemists in recent years. In addition, quinazolinones possess stability and change adaptability that makes it easy to prepare, which is prime for scientific investigations. Peharmaline A is found in Peganum harmala L., a species coming from the family of medicinally important Zygophyllaceae. In Chinese traditional medicine, P. harmala L. seed extract is widely used to treat malignancies of the digestive system and malaria. 

(±}-Peharmaline A (1} 

Figure 1: (±)-Peharmaline A precursors 

Objective of the Proposal 

We proposed to develop a novel approach to synthesize Peharmaline A utilizing the deoxyvasicinone and 5-methoxytryptamine, and extend the same to access a library of analogues of peharmaline to further study their biomedicinal potential. 

Methodology for the Synthesis of (±)-Peharmaline A and Its Analogues 

According to our retrosynthetic plan shown in Scheme 1, we have proposed a total synthetic route to achieve (±)-Peharmaline A 1 in a three-step process: through Pictet-Spengler reaction with the use of 5- methoxytryptamine 3 and deoxyvasicinone methyl oxalate intermediate 2, which itself can be made by of the acylation reaction of deoxyvasicinone 4 with methyl oxalyl chloride 5. It required us to develop a short and effective synthesis for deoxyvasicinone itself through adapting and modifying the reported synthesis into a practical largescale under microwave irradiation chemistry one-pot synthesis. Starting from commercially available isatoic anhydride 6 and pyrrolidinone 7, they would serve as the building blocks and will result in (±)-Peharmaline in just three steps. It would also provide a divergent approach towards the (±)-Peharmaline A analogues in a modular fashion. 

Scheme 1: Retrosynthetic plan of synthesizing (±)-Peharmaline A 

Results and Discussions 

Scheme 1 describes our efforts toward the development of a total synthetic approach to the synthesis of (±) -Peharmaline A. After a lot of careful experimentation with stoichiometry and heating (temperatures and time intervals), we finally optimized the synthesis of deoxyvasicinone to work on 1Og scale. It required thorough mixing of 1:1.25 equals of isatoic anhydride 5 and the pyrrolidinone 6 and heating of three minutes at medium power household MW (with one-minute intervals). Once the reaction was completed, it was allowed to cool to room temperature, and any remaining starting materials were removed under reduced pressure. The next step was acylation of the deoxyvasicinone 4. The acylation reaction conditions needed a lot of optimizations. We started with the available chemical in the lab i.e., ethyl oxalyl chloride 6 instead of methyl oxalyl chloride, which will lead to Methyl (±)-Peharmaline A, while we awaited the other chemical to arrive from the vendor. 

After a great deal of experimentation to optimize the temperature conditions, the equivalence of acylation reaction, and the solvent environment, we found out that 3 eq of ethyl oxalyl chloride and 2 eq of Et3N base in anhydrous DCM was needed to complete the reaction with deoxyvasicinone, resulting in  32.6 % of yield. Further, it was confirmed that the initially maintained O °C temperature needs to be increased after the complete addition (1 h) of ethyl oxalyl chloride to overnight refluxing to 48 h for the completion of the reaction to obtain a green clean solid product. The addition of both ethyl oxalyl chloride and Et3N base was separated into two portions, with the first half equivalent being added in at OoC initially, and the second half equivalent being added after 24 hours of refluxing in the same method before placing the reaction back to reflux for an additional 24 h. The acylated deoxyvasicinone 2 was obtained in enol form and it was confirmed by NMR and X-ray crystallography. 7 The methyl oxalyl chloride had arrived from the vendor and was tested with the same chemical and temperature environment, and experiments are still on-going to examine the identity and quality of the product. Additional tests are being conducted with deoxyvasicinone with benzoyl chloride and oxalyl chloride to determine if analogues can be developed to further the research with the final (±)-Peharmaline A and other quinazolinone products. 

Scheme 2: Synthesis approach of (±)-Peharmaline A analogue 

 

The next step in the sequence was the Pictet-Spengler reaction. We decided to establish it first with the commercially available tryptamine instead of the expensive 5-methoxy tryptamine (required to be synthesized in the lab later). Accordingly, we combined the tryptamine base 3.85b with acylated deoxyvasicinone 2 under refluxing conditions. After a lot of experimentation trials, our efforts towards the Pictet-Spengler reaction conditions were successful. It ensued in the presence of a catalytic amount of acid that could facilitate the reaction forward to the formation of (±)-Peharmaline A analogue when there was no water present in the sample. As we ran these optimization reactions on a mg scale, we obtained the product through preparative TLC thus far. We need to further optimize them on a larger scale and obtain the pure compound. 

 

Scheme 3: Pictet-Spengler reaction for the final assembly 

 

Figure 2 shows the changes in the 1H NMR peaks along the path of (±)-Peharmaline A analogue (8) (C) formation from the deoxyvasicinone 4 (A). 

Figure 2: Comparison of (±)-Peharmaline A analogue synthesis process A) 1H NMR of Deoxyvasicinone 4 in CDCl3 synthesized through MW irradiation chemistry. B) 1H NMR of enol form of acylated deoxyvasicinone 2 in CDCl3) 1H NMR of Peharmaline A analogue 8 in CDCl3 synthesized in TFMS acid/DCM solution. 

Conclusions and Future Directions

In summary, we have established the groundwork needed for the synthesis of peharmaline A and its analogues. Scheme 4 deliniates our approach of making a library of (±)-Peharmaline A analogues by employing different tryptamines and tyrosinesbased primary arylethanamines in Pictet-Spengler reaction by applying our optimized reaction conditions (anh. DCM/TFMSA/reflux-24 h) towards appending the b-Carboline ring to the deoxyvasicinone ring. As can be seen, it is a modular and divergent approach to making (±)-Peharmaline A analogues via employing different acid chlorides as well, as seen in Scheme 5. We propose to carry it out when the grad student that I was associated with returns in the next winter and summer breaks. 

 

Scheme 4: Future study: Scoping Pictet-Spengler towards (±)-Peharmaline A analogues 

Scheme 5: Other modifications for Diversity Oriented Synthesis of Peharmaline A 

Acknowledgments

Support from the UMassD OUR is greatly appreciated. Many thanks to Fazmina Anver and Dr. Rasapalli for their teaching and mentoring in the lab. 

References

Kulkarni, A. S.; Dash, A.; Shingare, R. D.; Chand, J.; Manhas, D.; Singh, A.; Nandi, U.; Goswami, A.; Srinivasa Reddy, D. Identification of New Modulator of DNA Repairing Pathways Based on Natural Product   (±)-Peharmaline   A.   Bioorg.   Med.   Chem.   2023,   91,   117365. https://doi.org/1O.1O16/j.bmc.2O23.117365. 

Piemontesi, C.; Wang, Q.; Zhu, J. Enantioselective Synthesis of (+)-Peganumine A. J. Am. Chem. Soc. 2016, 138 (35), 11148-11151. https://doi.org/1O.1O21/jacs.6bO7846. 

K.-B. Wang, S.-G. Li, X.-Y. Huang, et al. (±)-Peharmaline A: a pair of rare B-carboline-vasicinone hybrid alkaloid enantiomers from Peganum harmalaEur J Org Chem, 2O17 (2O17), pp. 1876-1879, 1O.1OO2/ejoc.2O17OO137 

Anver, F; Rasapalli, S. Thesis: Synthetic Studies Towards Biologically Active Heterocyclic Alkaloids and Their Analogues a dissertation in Chemistry and Biochemistry 

Alsibaee, A. M., Al-Yousef, H. M., & Al-Salem, H. S. (2O23). Quinazolinones, the Winning Horse in Drug Discovery. Molecules (Basel, Switzerland), 28(3), 978. https://doi.org/1O.339O/molecules28O3O978 

Cao, R.; Peng, W.; Wang, Z.; Xu, A. B-Carboline Alkaloids: Biochemical and Pharmacological Functions. Curr. Med. Chem. 2007, 14, 479-5OO. https://doi.org/1O.2174/O929867O777994O998. 

Alshehry, R; Rasapalli, S. Synthetic studies toward biologically active quinazolinones : a dissertation in Chemistry and Biochemistry https://umassd.primo.exlibrisgroup.com. 

 

 

Research in Biology

Human-Induced Fear in Free-Living Raccoons 

By Ruby Sanger 
Abstract 

This past June and July, I spent time at Norcross Wildlife Sanctuary in Wales, Massachusetts. Norcross is a large intact forest of approximately 4200 ha with restricted human access and a small 17ha area near the visitor center, where human access is permitted for hiking along 4km of natural trails. 

I intended to study the effects of human-induced fear on the foraging effects of free-living white-tailed deer, both in the human-accessible area and the forest. This design included audio treatments (human talking and birdsong) and food treatments (plain corn and plain corn mixed with molasses) at eight different stations across the sanctuary. However, after a few weeks of experimentation, there was no deer activity at any of the stations. While thorough background research was conducted to decide the most appealing foods to bait deer with, and there was known to be a high density of deer at the sanctuary, it is possible that because the forest is so dense with natural browse and plants the deer eat, they simply not interested in what was being offered. As a result, Dr. Sherriff and I decided to adapt the experiment to study the foraging behaviors of raccoons, as there were reoccurring raccoon visits at select feeding stations on the public trails. 

None of the stations with molasses had traffic, nor did any of the stations in the deep woods, so we cut the molasses treatment while adding two audio treatments (human yelling and dog barking) for a total of four treatments. The experiment was limited to the public trails of the sanctuary. 

Introduction 

While predators can kill prey, they can also alter prey fitness through nonconsumptive effects. The risk of predation can subsequently alter prey’s behavior, morphology, and physiology, which may all impact prey survival and reproduction. These effects may then influence prey population sizes (Sheriff et al., 2020). For example, a study by Cherry et al. (2016) showed that the presence of coyotes led to reduced lactation and ovulation in white-tailed deer, and an absence of coyotes related to an increase in feeding, lactation, and ovulation. These non-consumptive effects may reduce prey population size due to a lack of reproductive success and fecundity (Say-Sallaz et al., 2019). 

The term “landscape of fear” is defined as the spatial variation in prey perception of predation risk. These “landscapes” combine the elements of the physical environment that prey may inhabit or forage in, the predation risk and how it varies across locations, and a prey’s response to predation risk. There are generally two methods of prey response, one being avoiding areas perceived as higher risk, and the other being changes in behaviors while in the areas perceived as higher risk. The perception and fear of predation may be able to drive community-level changes within ecosystems, such as trophic cascades. (Gaynor et al., 2019). A common way to study the spatial variation resulting from risk response is by looking at giving-up densities (GUDs) and analyzing foraging behaviors within the context of risk. GUDs are used to provide insight into metabolic and predation costs of foraging by determining when an individual may stop foraging (Brown, 1987). 

In a study by Darimont et al. (2015), it has been shown that human predators kill far more prey than non-human predators, as well as killing carnivores nine times more than natural predators (Smith et al. 2017). While humans kill at an unparalleled rate, they more often affect prey behavior through disturbance (Frid, Dill 2002). Fear of humans as a “super predator” is also known to lead to behavioral changes in both predators and prey, and the subsequent effects on populations and communities may be larger than those resulting from non-human predators (Crawford et al. 2022). Experimental non-consumptive behaviors from humans have even led to a decrease in feeding times for pumas, an animal without any natural predators (Smith et al. 2017). A landscape of fear of the perception of humans can result in significant changes in wildlife behavior and community dynamics. Suraci et al. (2019) conducted studies with free-living mountain lions, bobcats, medium-sized carnivores (such as opossums and skunks) and deer mice in the Santa Cruz mountains. 

Human predation risk was simulated by using playbacks of human vocalization. The carnivore groups all experienced behavioral changes in response to perceived predation risk: avoiding areas, making temporal changes, and being less efficient in foraging. 

However, deer mice seemingly benefited from human presence; they increased space use as well as foraging intensity. The fear that the carnivores perceived affected lower trophic levels, influencing the surrounding wildlife system (Suraci et al., 2019).

Methods 

i. Feeding Station Set-ups 

Figure 1: Map of the 8 feeding stations, marked on the GAIA app. 

 

Figure 2: Camera and speaker set-ups. The cups were used as a shield for the exposed speakers from the rain. 

 

Eight feeding stations were chosen among the public trails of Norcross Wildlife Sanctuary in Wales, Massachusetts (Figure 1). Each station consisted of a painted feeding bin, a field camera, and a speaker (Figure 2). 

The speaker and camera were programmed to be used together, using Arduino, so that when the camera was triggered by motion, the speaker was triggered to play a programmed playback. Two speakers were programmed to play conversational human speaking, two to play dogs barking, two to play bird songs native to the area, and two to play humans yelling. Each speaker was programmed to play at about 65-70 decibels. The camera-speaker set-ups were programmed for the speaker to trigger 20 seconds after the camera was triggered. 

80 ounces of dried whole grain corn (5 lbs) were set into the feeding tubs. Each empty tub, when closed with the lid, weighed 70 ounces. 

ii. Daily Protocol 

The stations were filled with 80 ounces of corn, and the cameras and speakers were switched on June 30th. Every morning from July 1st to July 10th, the combined weight of the corn and the tub were taken at each of the eight locations by closing the tub and weighing it with a digital fishing scale. If the weight was below 120 ounces, the corn was later refilled to the base weight of 150 ounces (80 ounces of corn plus the weight of the tub). Additionally, the SD cards in the cameras were checked to see if there was any raccoon activity or other significant animal activity during the night. The battery levels of the cameras and speakers, as well as the SD card storage amounts, were also checked every morning to ensure proper performance for the following night. In the case of heavy rain, the tubs were covered to prevent the corn from being waterlogged, which could provide inaccurate weights. 

Table 1: Key of auditory playbacks per feeding station

Feeding Station  Treatment 6/30-7/05  Treatment 7/05-7/10 
1  Talking  Birdsong 
2  Yelling  Dog barking 
3  Dog barking  Yelling 
4  Birdsong  Talking 
5  Talking  Birdsong 
6  Yelling  Dog barking 
7  Birdsong  Yelling 
8  Dog barking  Talking 

On July 5th, after 5 nights of data collection, the playback treatments were changed to different locations (Table 1). At the end of data collection, the SD cards were collected, and the feeding stations were broken down. 

iii. Data Analysis 

The footage captured from the feeding stations are currently being analyzed manually. I will be scoring for behaviors including fleeing (running/leaving quickly), leaving (walking away), looking up, head-up foraging, head-down foraging, playing, freezing, and leaving a group. There are many cases of corn being eaten by mice, chipmunks, squirrels, and occasionally deer. The footage of these animals will be used to separate their consumption from the raccoon’s consumption. Additionally, statistics of each night of the experimental run are being gathered. I am counting the total number of foraging events as well as the time spent eating each night. The footage is currently being analyzed. 

Conclusion 

While the results are still being analyzed, information regarding how raccoons react to auditory playbacks will provide useful insight into the effectiveness of using sound as a method of pest control, as well as how small mammals such as raccoons are affected by the presence of humans. I hope to continue in this line of study and resume with the original design for studying white-tailed deer in the future. 

Works Cited 

Sheriff MJ, Peacor SD, Hawlena D, Thaker M. (2020). “Non‐consumptive predator effects on prey population size: A dearth of evidence.” Journal of Animal Ecology vol 89. 

Cherry, M. J., K. E. Morgan, B. T. Rutledge, L. M. Conner, and R. J. Warren. (2016). “Can coyote predation risk induce reproduction suppression in white-tailed deer?” Ecosphere 7(10):01481. 

Say-Sallaz, E., Chamaille-Jammes, S., Fritz, H., Valeix, M. (2019). “Non-consumptive effects of predation in large terrestrial mammals: Mapping our knowledge and revealing the tip of the iceberg.” Biological Conservation vol.235: 46-52. 

Suraci, J.P, Clinchy, M., Zanette, L.Y., Wilmers, C.C. (2019). “Fear of humans as apex predators has landscape-scale impacts from mountain lions to mice.” Ecology letters vol. 22,10: 1578-1586. 

Gaynor, K.M., Brown, J.S., Middleton, A.D., Power, M.E., Brashares, J.S. (2019). “Landscapes of Fear: Spatial Patterns of Risk Perception and Response.” Trends in ecology & evolution vol. 34,4: 355-368. 

Darimont, C.T, Fox, C.H, Bryan, H.M, Reimchen, T.E. (2015). “HUMAN IMPACTS. The unique ecology of human predators.” Science vol. 349,6250: 858-60. 

Smith J.A., Suraci J.P, Clinchy M., Crawford A., Roberts D., Zanette L.Y., Wilmers C.C. (2017). “Fear of the human ‘super predator’ reduces feeding time in large carnivores.” Proceedings. Biological sciences vol. 284,1857: 20170433. 

Frid, Alejandro, and Dill, L. (2002). “Human-Caused Disturbance Stimuli as a Form of 

Predation Risk.” Conservation Ecology, vol. 6. 

Crawford, D.A., Conner, M.L, Clinchy, M., Zanette, L.Y., Cherry, M.J. (2022). “Prey tells, large herbivores fear the human ‘super predator’.” Oecologia vol. 198,1: 91-98. 

Gaynor K., Hojnowski C., Carter N., Brashares J. (2018). “The influence of human disturbance on wildlife nocturnality.” Science vol. 360,6394 (2018): 1232-1235. 

1 2 3 9