ArticlePDF Available

Bladder cancer therapy using a conformationally fluid tumoricidal peptide complex

Authors:

Abstract and Figures

Partially unfolded alpha-lactalbumin forms the oleic acid complex HAMLET, with potent tumoricidal activity. Here we define a peptide-based molecular approach for targeting and killing tumor cells, and evidence of its clinical potential (ClinicalTrials.gov NCT03560479). A 39-residue alpha-helical peptide from alpha-lactalbumin is shown to gain lethality for tumor cells by forming oleic acid complexes (alpha1-oleate). Nuclear magnetic resonance measurements and computational simulations reveal a lipid core surrounded by conformationally fluid, alpha-helical peptide motifs. In a single center, placebo controlled, double blinded Phase I/II interventional clinical trial of non-muscle invasive bladder cancer, all primary end points of safety and efficacy of alpha1-oleate treatment are reached, as evaluated in an interim analysis. Intra-vesical instillations of alpha1-oleate triggers massive shedding of tumor cells and the tumor size is reduced but no drug-related side effects are detected (primary endpoints). Shed cells contain alpha1-oleate, treated tumors show evidence of apoptosis and the expression of cancer-related genes is inhibited (secondary endpoints). The results are especially encouraging for bladder cancer, where therapeutic failures and high recurrence rates create a great, unmet medical need.
Tumoricidal activity of two non-homologous alpha-helical peptide–oleate complexes a Ribbon representation of the crystallographically determined three-dimensional structure of human α-lactalbumin (PDB ID: 1B9O), indicating the alpha1 (blue), beta (green), and alpha2 (gray) domains. The calcium ion is not shown. b Far-UV circular dichroism spectra of synthetic alpha1 peptide, beta peptide, and their respective peptide–oleate complexes. c, d Death response in human lung (A549), kidney (A498), and murine bladder (MB49) carcinoma cells, quantified as a reduction in ATP levels (c, P = 3.26E−5 for A549, 0.013 for A498 and 0.005 for MB49, alpha1–oleate compared to beta–oleate) or PrestoBlue fluorescence (d, P = 0.007 for A549, 0.003 for A498 and 0.002 for MB49, alpha1–oleate compared to beta–oleate). Cells were treated with the alpha1–oleate complex (blue) or the beta–oleate complex (green), (3 h, 35 μM, cell death compared to PBS controls). For controls exposed to the naked peptides or oleate alone, see Supplementary Fig. 1d. e Colony assay showing dose-dependent long-term effects of alpha1-oleate. A representative image is shown from two independent experiments. Scale bar = 5 mm. f Alpha1–oleate triggers rapid membrane blebbing in A549 lung carcinoma cells (35 μM, 10 min). Scale bar = 10 μm. A representative image is shown from three independent experiments. g K⁺ efflux in A549 lung carcinoma cells exposed to alpha1–oleate and inhibition with BaCl2. h Inhibition of cell death by the ion flux inhibitors Amiloride and BaCl2 (100 μM), measured by PrestoBlue fluorescence (P = 0.031 for 21 μM + BaCl2, 0.005 for 21 μM + Amiloride, 0.028 for 35 μM + BaCl2, and 0.014 for 35 μM + Amiloride, compared to no inhibitor). i DNA strand breaks detected by TUNEL staining in alpha1–oleate-treated A549 lung carcinoma cells (n = 50 cells per group). Scale bar = 20 μm. j AlexaFluor568-labeled alpha1–oleate (red) is internalized by A549 lung carcinoma cells. Nuclei are counterstained with DAPI (blue) (n = 52 cells per group). Scale bar = 10 μm. Data are presented as mean ± SEM from three independent experiments, *P < 0.05, **P < 0.01, ***P < 0.001, analyzed by two-tailed unpaired t-test (c, d, h, j) and 2-way ANOVA using Dunnett’s correction (i).
… 
Primary endpoints: shedding of tumor cells and reduction in tumor size following intra-vesical instillation of alpha1–oleate a–c Cell shedding increased significantly after alpha1–oleate instillation. a Scatterplot showing individual means of six visits per patient in the treatment group (n = 20) compared to patients receiving placebo (n = 20). Line represents the median. b Comparison of cell numbers in urine before (pre = white) and after (post = black) alpha1–oleate inoculation on visits 1–6 showing increased cell numbers post-inoculation in the treatment group (n = 20 patients per group, P = 0.0030 for visit 1, 0.0098 for visit 2, <0.0001 for visits 3 and 4, 0.0073 for visit 5 and 0.0336 for visit 6) but not in the placebo group. Data are presented as mean ± SEM. c Representative images, illustrating the increase in cell shedding after alpha1–oleate instillation. Magnification = ×400. Scale bar = 50 μm. d–f Difference in the shedding of tumor cell clusters between the treatment and placebo group. d Scatterplot showing individual means of six visits per patient in the treatment group compared to patients receiving placebo. Line represents the median. e Increased numbers of cell clusters in post-inoculation samples of patients receiving alpha1–oleate (n = 20 patients per group, P = 0.9743 for visit 1, 0.0212 for visit 2, <0.0001 for visits 3, 4, and 5, and 0.0005 for visit 6). Data are presented as mean ± SEM. f Representative images of cancer cell clusters after alpha1–oleate instillation. Magnification = ×400. Scale bar = 50 μm. g Paris grade of shed cells before or after alpha1–oleate instillation. An increase is observed in the treatment group (χ² test). h Reduction in tumor size in patients receiving alpha1–oleate treatment. Images were compared between the time of diagnosis and the time of TURB (P = 0.04, χ² test for trend compared to placebo, n = 19 for treatment group and n = 20 for placebo group). i Examples of cystoscopy photographs obtained by A.B. at the time of diagnosis and after treatment at the time of TURB. Scale bars = 5 mm. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001. The data were analyzed by two-tailed unpaired Mann–Whitney U-test (a, d) or by repeated-measures two-way ANOVA with Sidak’s correction (b, e).
… 
This content is subject to copyright. Terms and conditions apply.
ARTICLE
Bladder cancer therapy using a conformationally
uid tumoricidal peptide complex
Antonín Brisuda1,7, James C. S. Ho 2,6,7, Pancham S. Kandiyal3, Justin T-Y. Ng4, Ines Ambite 2,
Daniel S. C. Butler 2, Jaromir Háček5, Murphy Lam Yim Wan2, Thi Hien Tran2, Aftab Nadeem2,
Tuan Hiep Tran2, Anna Hastings 3, Petter Storm2, Daniel L. Fortunati3, Parisa Esmaeili2, Hana Novotna1,
Jakub Horňák1,Y.G.Mu 4, K. H. Mok 3, Marek Babjuk1,8 & Catharina Svanborg 2,8
Partially unfolded alpha-lactalbumin forms the oleic acid complex HAMLET, with potent
tumoricidal activity. Here we dene a peptide-based molecular approach for targeting and
killing tumor cells, and evidence of its clinical potential (ClinicalTrials.gov NCT03560479). A
39-residue alpha-helical peptide from alpha-lactalbumin is shown to gain lethality for tumor
cells by forming oleic acid complexes (alpha1-oleate). Nuclear magnetic resonance mea-
surements and computational simulations reveal a lipid core surrounded by conformationally
uid, alpha-helical peptide motifs. In a single center, placebo controlled, double blinded Phase
I/II interventional clinical trial of non-muscle invasive bladder cancer, all primary end points
of safety and efcacy of alpha1-oleate treatment are reached, as evaluated in an interim
analysis. Intra-vesical instillations of alpha1-oleate triggers massive shedding of tumor cells
and the tumor size is reduced but no drug-related side effects are detected (primary end-
points). Shed cells contain alpha1-oleate, treated tumors show evidence of apoptosis and the
expression of cancer-related genes is inhibited (secondary endpoints). The results are
especially encouraging for bladder cancer, where therapeutic failures and high recurrence
rates create a great, unmet medical need.
https://doi.org/10.1038/s41467-021-23748-y OPEN
1Department of Urology, Motol University Hospital, 2nd Faculty of Medicine, Charles University Praha, Prague, Czech Republic. 2Division of Microbiology,
Immunology and Glycobiology, Department of Laboratory Medicine, Faculty of Medicine, Lund University, Lund, Sweden. 3Trinity Biomedical Sciences
Institute (TBSI), School of Biochemistry & Immunology, Trinity College Dublin, The University of Dublin, Dublin, Ireland. 4School of Biological Sciences,
Nanyang Technological University, Singapore, Singapore. 5Department of Pathology and Molecular Medicine, Motol University Hospital, 2nd Faculty of
Medicine, Charles University Praha, Prague, Czech Republic.
6
Present address: Centre for Biomimetic Sensor Science, Nanyang Technological University,
Singapore, Singapore.
7
These authors contributed equally: Antonín Brisuda, James C.S. Ho.
8
These authors jointly supervised this work: Marek Babjuk,
Catharina Svanborg. email: catharina.svanborg@med.lu.se
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 1
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Targeted cancer therapies have made signicant advances
but the lack of tumor specicity remains a signicant
concern1,2. Few current therapies kill cancer cells without
harming healthy tissues, and severe side effects have become
accepted as a necessary price to pay for survival or cure. The
notion of successfully combining efcacy with increased tumor
selectivity is justly regarded with skepticism. Yet, a serendipitous
discovery has provided insights into mechanisms of tumor-
specic cell death, induced by unfolded polypeptide chains, which
acquire tumoricidal activity by forming fatty acid complexes3,4.
Extensive observations in tumor models and clinical studies have
further dened the proteinlipid complexes as an interesting class
of molecules with signicant therapeutic potential5.
The ndings are challenging, as protein unfolding and loss of
structural denition is associated with a gain of toxicity, due to
the formation of amorphous aggregates and amyloid brils6,7.
Native protein structure is often regarded as a prerequisite for
biological function, by epitope-specic interactions and molecular
tness for a nite number of cellular targets. Yet, a lack of
structural denition may, in some cases, result in a gain of
function, in part by uncovering different conformations and
exposing peptide motifs that are unavailable in the native state8,9.
Such effects have been predicted for membrane perturbing α-
helices in antimicrobial peptides, where the ability to destabilize
lipid bilayers has been proposed to reside in the three-
dimensional conformation rather than the amino acid
sequence10.
Alpha-lactalbumin is crucial for the survival of lactating
mammals. In its native state, the protein serves as a substrate
specier in the lactose-synthase complex11,dening the nutri-
tional content and uidity of milk. Partially unfolded alpha-lac-
talbumin, in contrast, forms an oleic acid complex, named
HAMLET, with potent tumoricidal activity3,4,8,9. The HAMLET
complex kills a range of tumor cells with rapid kinetics and shows
therapeutic efcacy in animal models of colon cancer, glio-
blastoma, and bladder cancer1215. Early, investigator-driven
clinical studies demonstrated that HAMLET is active topically,
against skin papilloma and induces tumor cell shedding into the
urine in patients with bladder cancer5,16.
This study presents a synthetic, peptide-based drug candidate
derived from alpha-lactalbumin, which reproduces the tumor-
icidal properties of HAMLET and allows for a full translation of
these ndings into the clinic. Through complementary nuclear
magnetic resonance (NMR) analysis and computational model-
ing, the molecular basis for this gain-of-functionis dened,
including the three-dimensional structural motifs that determine
fatty acid-binding efciency and tumoricidal activity. The ther-
apeutic efcacy of the complex is demonstrated in patients with
non-muscle invasive bladder cancer (NMIBC), in a fully con-
trolled clinical trial.
Results
Peptide-specic tumoricidal activity. To understand the invol-
vement of specic peptide motifs in tumor cell death, we syn-
thesized the N-terminal alpha-helical domain (residues 139,
alpha1) or the beta-sheet (4080, beta) domains of human alpha-
lactalbumin (Fig. 1a). The alpha1 peptide formed complexes with
oleate (alpha1oleate, 1:5) and circular dichroism (CD) spectra
detected an increase in alpha-helical structure content in these
complexes (Fig. 1b). The betaoleate complex remained structu-
rally unchanged (Fig. 1b). Alpha1oleate triggered a rapid, dose-
dependent death response in human lung- and kidney carcinoma
cells and in murine bladder cancer cells (Fig. 1c, d). The
betaoleate complex lacked tumoricidal activity and tumor cells
subjected to the naked alpha-helical peptides (35 μM) or oleate
(175 μM) controls were not tumoricidal (Fig. 1c, d and Supple-
mentary Fig. 1). The loss of cell viability was irreversible, as
shown after 10 days, by using colony assays (Fig. 1e and Sup-
plementary Fig. 1). Membrane blebbing occurred in tumor cells
within minutes of exposure to alpha1oleate but the naked
peptide- and oleate controls were not active (Fig. 1f and Sup-
plementary Fig. 1). Rapid K+uxes were recorded, further
dening the membrane response (Fig. 1g). Pretreatment of the
cells with Na+and K+ux inhibitors reduced cell death by
4050%, linking the membrane response to tumor cell death
(Fig. 1h). The alpha1oleate complex was rapidly internalized by
tumor cells and by TUNEL staining, alpha1oleate was shown to
induce double-strand DNA breaks in the tumor cells, indicative
of apoptosis (Fig. 1i, j).
In a screen of proteins with membrane-integrating properties,
SAR1 was found to form tumoricidal complexes with oleic acid,
reproducing effects of alpha1oleate (Supplementary Figs. 1 and
2). SAR1 is a membrane-integrating protein of the COPII
complex that induces membrane tubulation by insertion of its N-
terminal amphipathic α-helix1719. The N-terminal alpha-helical
peptide (residues 123, sar1alpha) formed an oleate complex,
which efciently killed tumor cells (Supplementary Figs. 1 and 2).
The naked peptide- and oleate controls were not active
(Supplementary Fig. 1). Sar1alphaoleate triggered membrane
blebbing in tumor cells, rapid K+uxes were recorded, and
tumor cell death was partially inhibited by Na+and K+ux
inhibitors, suggesting a similar mode of action of the two
complexes, despite low sequence homology (Supplementary
Fig. 2). The sar1betaoleate complexes and naked peptide
controls did not trigger tumor cell death, however (Supplemen-
tary Figs. 1 and 2).
In preparation for the clinical trial, the safety of alpha1oleate
was investigated in C57BL/6J mice carrying MB49-induced
bladder tumors13. Therapeutic efcacy in 100% of treated mice
compared to the sham group and a lack of toxicity was
demonstrated, providing the necessary background to plan the
clinical trials13.
Biomolecular NMR analysis of the peptideoleate complexes.
1H NMR spectra of the alpha1oleate and sar1alphaoleate
complexes detected a shift from sharp signals for the naked
peptides to broad signals and poor chemical shift dispersion for
the oleate complexes (Fig. 2a, b), suggesting a conformational
change from a random-coil fast-exchange time regime to an
intermediate millisecond timescale. Broadening in the amide,
side-chain methyl and aromatic regions suggests that interactions
between fatty acids and peptides occur throughout the molecules.
Two-dimensional nuclear Overhauser effect spectroscopy (2D
NOESY) spectra identied non-covalent, relatively short through-
space interactions between the respective peptides and fatty acids.
Important nuclear Overhauser effects (NOEs) were detected
between the olenic protons (5.23 ppm) of oleic acid and the Hα
and aromatic protons of alpha1 and between the sar1alpha aro-
matic region and the oleic acid olenic protons (Fig. 2c, d). The
downeld chemical shift of amide protons observed between 7.6
and 8.8 ppm suggests the presence of secondary structure in
alpha1 and alpha1oleate. Well-resolved signals obtained from
the one-dimensional 1H NMR spectra provided a stoichiometry
of 3.7 oleate molecules per alpha1 peptide. Chemical shift map-
ping revealed a cluster of residues with aliphatic side chains that
change upon the binding of oleate, providing further evidence of
interactions between peptides and fatty acids (Fig. 2e, f).
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
2NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Hydrodynamic volume measurements carried out with size-
exclusion high-performance liquid chromatography (SE-HPLC)
and diffusion-ordered NMR spectroscopy (DOSY) showed that
the alpha1oleate complex R
H
was considerably larger (27.4 and
29.3 Å, respectively) than the naked peptides (16.1 Å from SE-
HPLC and 14.4 Å from DOSY) (Fig. 2g, h and Supplementary
Figs. 3 and 4). The distinctly larger R
2
values (transverse
relaxation rate) for the complex than those of alpha1 peptide
and human serum albumin (HSA) suggested slower millisecond
to microsecond exchange processes (Supplementary Table 1 and
Supplementary Figs. 5 and 6). Importantly, the R
2
values for the
complex were also different from oleate in an aqueous solution,
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
suggesting that the dynamics of the complex were clearly different
from micelle/vesicle-like particles formed from oleic acid/oleate
with no peptide component.
Computational analyses of the peptide- and peptideoleate
system. Computational simulations also pointed toward struc-
tural heterogeneity, showing that the naked peptides and
peptideoleate complexes belonged to wide conformational
spaces with relatively deep basins (Fig. 3a, b). Representative
structures mapped to different free energy surface minima,
revealed prominent alpha-helical secondary structural elements
and a hydrophobic oleate core for the peptideoleate complexes
(Fig. 3c, e). The peptideoleates fold upon this core differently
from the naked peptides, which exhibit multiple local minima
(Fig. 3d, f, and Supplementary Tables 2 and 3). Naked alpha1
ensembles were characterized by various partially folded helix-
turn conformations, whereas naked sar1alpha ensembles exhib-
ited a mixture of the random coil, alpha-helical, and beta-sheet
structures.
A contact probability analysis revealed that the interactions
between alpha1 or sar1alpha and oleate were mainly hydro-
phobic, with a >0.9 contact probability with olenic protons
(Supplementary Tables 4 and 5). The peptideoleate complexes
displayed relatively wide and deep free energy minima basins,
suggesting that a multitude of conrmations would be equally
possible to visit (Fig. 3c, e). When combined with the R
2
relaxation rates, the possibility of multiple sampling of various
conformations within a short period of time provides an
argument that rather than targeting specic partners, these
alpha-helical complexes may potentially be interacting with
multiple putative binding partners available on cancer cell
surfaces20.
Based on these extensive investigations and the strong agreement
of the experimental aspects with the simulated predicted ensembles,
it was clear that the apparently unrelated peptides alpha1 and
sar1alpha can form complexes with shared structural characteristics,
involving a exible peptide moiety and a fatty acid cluster. This
notion resonates with the sequence-function inconsistency among
certain antimicrobial amphipathic alpha helices, where peptides
with similar overall properties, such as hydrophobicity or charge,
can have dramatically different levels of activity10.
A placebo-controlled, randomized clinical trial of alpha1oleate
in patients with NMIBC. NMIBC is common and despite current
treatment protocols, recurrence rates are high21,22. To address if
the therapeutic effects observed in the murine MB49 bladder
cancer model can be translated into the clinic, the investigational
product alpha1oleate was produced under GMP conditions. The
alpha1oleate complex was further subjected to formal toxicity
testing and the results have been published13. Toxicity for bladder
tissue was not detected at concentrations ranging from 1.7 to 17
mM13.
The clinical safety and therapeutic potential of alpha1oleate
were tested in a single-center, placebo-controlled, double-blind
Phase I/II trial (EudraCT No: 2016-004269-14, ClinicalTrials.gov
NCT03560479, Supplementary Note 1, Supplementary Table 6).
Patients with suspected NMIBC were randomized 1/1 to receive
alpha1oleate or placebo during a period of 22 days, prior to
endoscopic removal of the tumor by transurethral resection
(TURB), (Fig. 4a, b). Alpha1oleate (1.7 mM) or placebo (PBS)
was administered intravesical on six occasions (30 mL, days 1, 3,
5, 8, 15, and 22). The placebo solution was identical in appearance
to the active treatment. Demographic data, medical history, and
vital signs did not differ between the treatment and placebo
groups (for details see Supplementary Table 7).
Primary study endpoints. Adverse events (AEs) were recorded
and coded according to MedDRA (version 21.1) with a safety
follow-up after 52 days (Supplementary Table 8). Procedure-
related AEs, such as dysuria and bacteriuria, occurred at a similar
rate in the treatment and placebo groups. AEs specic for the
treatment group were not detected, suggesting low toxicity of the
study medication (Fig. 4c). Furthermore, there was no evidence of
a toxic response in healthy tissue samples from patients treated
with alpha1oleate, dened by histopathology or TUNEL
staining.
Tumor cell shedding and release of tumor cell clusters were
recorded. Cells with uroepithelial morphology were quantied in
urine at each visit, before instillation and about 2 h after the
instillation of alpha1oleate or placebo. Alpha1oleate triggered a
rapid increase in cell shedding compared to the pre-instillation
sample in all treated patients, at all visits (Fig. 5ac and
Supplementary Fig. 7). In addition, tumor cell clusters were
released into the urine in the treatment group. The clusters were
relatively large and the presence of tumor stroma in some samples
supported their tumor origin (Fig. 5df). The cells shed in urine
were assigned a pathology score as per the Paris classication
(classes 16, urine cytology was a secondary endpoint). In the
treatment group, an increase in the Paris score was detected in
post-inoculation samples compared to pre-inoculation samples
(Fig. 5g). A low level of cell shedding in the placebo group was
attributed to the instillation procedure and changes in pathology
Fig. 1 Tumoricidal activity of two non-homologous alpha-helical peptideoleate complexes. a Ribbon representation of the crystallographically
determined three-dimensional structure of human α-lactalbumin (PDB ID: 1B9O), indicating the alpha1 (blue), beta (green), and alpha2 (gray) domains.
The calcium ion is not shown. bFar-UV circular dichroism spectra of synthetic alpha1 peptide, beta peptide, and their respective peptideoleate complexes.
c,dDeath response in human lung (A549), kidney (A498), and murine bladder (MB49) carcinoma cells, quantied as a reduction in ATP levels (c,P=
3.26E5 for A549, 0.013 for A498 and 0.005 for MB49, alpha1oleate compared to betaoleate) or PrestoBlue uorescence (d,P=0.007 for A549,
0.003 for A498 and 0.002 for MB49, alpha1oleate compared to betaoleate). Cells were treated with the alpha1oleate complex (blue) or the betaoleate
complex (green), (3 h, 35 μM, cell death compared to PBS controls). For controls exposed to the naked peptides or oleate alone, see Supplementary Fig. 1d.
eColony assay showing dose-dependent long-term effects of alpha1-oleate. A representative image is shown from two independent experiments. Scale
bar =5 mm. fAlpha1oleate triggers rapid membrane blebbing in A549 lung carcinoma cells (35 μM, 10 min). Scale bar =10 μm. A representative image is
shown from three independent experiments. gK+efux in A549 lung carcinoma cells exposed to alpha1oleate and inhibition with BaCl
2
.hInhibition of
cell death by the ion ux inhibitors Amiloride and BaCl
2
(100 μM), measured by PrestoBlue uorescence (P=0.031 for 21 μM+BaCl
2
, 0.005 for 21 μM+
Amiloride, 0.028 for 35 μM+BaCl
2,
and 0.014 for 35 μM+Amiloride, compared to no inhibitor). iDNA strand breaks detected by TUNEL staining in
alpha1oleate-treated A549 lung carcinoma cells (n=50 cells per group). Scale bar =20 μm. jAlexaFluor568-labeled alpha1oleate (red) is internalized
by A549 lung carcinoma cells. Nuclei are counterstained with DAPI (blue) (n=52 cells per group). Scale bar =10 μm. Data are presented as mean ± SEM
from three independent experiments, *P< 0.05, **P< 0.01, ***P< 0.001, analyzed by two-tailed unpaired t-test (c,d,h,j) and 2-way ANOVA using
Dunnetts correction (i).
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
4NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
score or shedding of cell clusters were not observed (Fig. 5ag and
Supplementary Fig. 7).
The endoscopic appearance of the tumors was recorded at the
time of diagnosis and prior to surgery using a exible cystoscope
with white light-band and narrow-band imaging. Sizes were
assessed by an experienced endourologist using fully opened
clamps of the exible forceps and a measuring device close to the
tumor. Paired images from 39 patients were evaluated in a
blinded manner, by an independent NMIBC expert using a
simplied Delphi method23 addressing changes in lesion size,
supercial necrosis, and tissue vascularization. A reduction in
lesion size was detected in the treatment group (n=19, Fig. 5h, i).
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
In 1 patient, paired pre-treatment and post-treatment images
could not be collected for technical reasons. No difference in
supercial necrosis or vascularization was observed and there was
no change in lesion size in the placebo group (n=20).
Secondary end points. Evidence of tumor cell apoptosis was
obtained by staining of tumor biopsies obtained at the time of
surgery. Biopsies were examined for evidence of treatment-
induced apoptosis, quantied by the TUNEL assay (Fig. 6ac and
Supplementary Fig. 8). A signicant increase in net mean uor-
escence was detected in the treatment group, compared to pla-
cebo (see also Fig. 1). Staining was most intense adjacent to the
lumen suggesting that a gradient might be formed, from the
lumen towards the center of the tumor. TUNEL staining intensity
was signicantly correlated to cell shedding and alpha1oleate
uptake in individual patients (Fig. 6d) but not to the tumor grade.
In healthy tissue biopsies from the treated patients, TUNEL
staining was low. Tumors from the placebo group did not show
increased TUNEL staining, suggesting that tumor cell apoptosis
may be induced by the alpha1oleate treatment (Fig. 6c).
The alpha1oleate content of shed cells in urine was quantied
by immunohistochemistry, using alpha1-specic antibodies.
Alpha1-staining was detected in 70% of post-inoculation samples
in the treatment group (Fig. 6e, f and Supplementary Fig. 9).
Uptake correlated with cell shedding and cluster grade but not
with the tumor grade or stage (Supplementary Fig. 9).
The response to alpha1oleate was further evaluated by RNA-
seq, using RNA from tumor biopsies and comparing the
treatment to the placebo group. A strong treatment effect was
detected (Fig. 7ac and Supplementary Fig. 10). Cancer-related
genes accounted for about 80% of the signicantly regulated
genes in the treated patients (cut off fold change > 2.0, P< 0.05),
conrming the effect of alpha1oleate on the tumor environment.
Genes regulating tumor growth and invasion were inhibited and
Ras signaling was suppressed, consistent with known effects of
the complex on Ras family members24 (Fig. 7d and Supplemen-
tary Fig. 10). Bladder cancer genes were specically regulated,
including metalloproteinases, solute carriers, WNT complex
constituents, and thrombospondin, which affects angiogenesis25
(Fig. 7e). Furthermore, Fatty Acid Desaturase 6 (FADS6) and
transcriptional activator CREB3L4 were affected, suggesting that
the tumors respond to the constituents of the alpha1oleate
complex. FADS6 regulates oleate biosynthesis and CREB3L4 the
unfolded protein response to conformationally uid proteins,
such as the alpha1 peptide. Interesting targets also included the
gap junction alpha1 protein, which was inhibited, potentially
promoting cell detachment (GJA1/CXA1 encoding Connexin 43,
Supplementary Fig. 10). No difference in tumor grade was
observed between the treatment or placebo groups, dened by
WHO 1973 and 2004/2016 criteria (Supplementary Table 9).
Data regarding two secondary end-points are not reported. As
this is an interim analysis, the long-term treatment effects will be
evaluated when the entire study has been completed. The urine
proteomics data set has not been fully analyzed.
Discussion
Bladder cancer is the fourth most common malignancy in the
United States and the fth in Europe, with a prevalence of about
1/400026. Due to high recurrence rates and a lack of curative
therapies, bladder cancer has the highest lifetime treatment costs
per patient of all cancers, followed by colorectal-, breast-, pros-
tate-, and lung cancer27. More than 80% recur after complete
surgical removal of the rst tumor and 15% progress to muscle-
invasive disease28. Intravesical chemotherapy and Bacillus
CalmetteGuérin (BCG) immunotherapy have limited efcacy
and signicant side effects29,30. Systemic administration of PD-1
and PD-L1 inhibitors is considered only in BCG unresponsive
patients where the experience is limited. Therapeutic options are
also limited by the inadequate supply of immunotherapy and
chemotherapy drugs worldwide31. In this study, we identify
conformationally uid peptidefatty acid complexes as additional
tools in cancer therapy and show that intra-vesical inoculation of
alpha1oleate is safe and effective in patients with bladder cancer.
The tumor response to alpha1oleate was analyzed in-depth,
using cellular and molecular tools to detect changes induced by
the complex. Treatment triggered the shedding of cells and tissue
fragments into the urine and alpha1oleate internalization by
tumor cells conrmed the afnity of the complex for the tumor.
Further analysis of tissue biopsies suggested a lasting effect of the
alpha1oleate instillations, as several tumor samples showed a
gradient-like pattern of apoptosis, starting from the bladder
lumen. Dysfunctional apoptosis has been identied as a key to
tumor development, especially in environments where oncogenes
such as MYC drive tumor cell proliferation32. Numerous attempts
have been made to develop apoptosis-inducing therapeutics with
tumor selectivity, but this has proven challenging, probably due to
the heterogeneity of individual tumors as well as their intrinsic
resistance to activating cell death pathways. The ability of
alpha1oleate to stimulate apoptosis in the majority of bladder
tumors is, therefore, encouraging and consistent with the
apparent lack of toxicity for bladder tissue.
RNA sequencing revealed profound molecular changes in
treated tissues, attributable to alpha1oleate. Classical cancer gene
networks were strongly inhibited in the treated patients, com-
pared to the placebo group, including Ras, previously identied as
a target for HAMLET; the oleate complex formed by the alpha-
lactalbumin holoprotein24. HAMLET binds activated Ras at the
plasma membrane of tumor cells and inhibits the Ras signaling
pathway, in part through effects on b-Raf phosphorylation. Sig-
nicant effects on adaptive immunity were not detected and
innate immunity was largely inhibited, including granulocyte
activation pathways. Notably, genes involved in oleate metabo-
lism and the unfolded protein response were affected, possibly
reecting a direct response to the constituents of the
alpha1oleate complex. In addition, treatment inhibited GJA1,a
Fig. 2 Biomolecular NMR analysis of naked alpha1- and sar1alpha peptides and their oleate complexes. a, b One-dimensional 1H NMR spectra. The
naked alpha1- (a, black) and sar1alpha- (b, black) peptides assume an ensemble of structures that interconvert rapidly and are therefore seen as sharp
peaks. The alpha1oleate (a, red) and sar1alphaoleate complexes (b, red) show broader peaks. Arrows indicate the indole 1H signals arising from the three
Trp side chains present in the sar1alpha peptide. c,dTwo-dimensional 1H NOESY spectra of alpha1oleate and sar1alphaoleate complexes, showing
atomic-level proximities of the fatty acid to the respective peptide. The spectra highlight NOEs between the 9,10 olenic protons (5.23 ppm) of oleic acid
with the Hαprotons and aromatic protons of the alpha1oleate complex (c) and the sar1alphaoleate complex (d). e,fTwo-dimensional 1H13C
Heteronuclear Single Quantum Coherence (HSQC) spectra overlays of the alpha1 peptide (red) and alpha1-oleate complex (black). Chemical shift
perturbation is detected in the aromatic side-chain region and the imidazole ring protons (e, green circled regions), and in the aliphatic side chain regions
(f). gSize-exclusion HPLC (SE-HPLC) of the alpha1 peptide and the alpha1oleate complex, mapped onto a standard calibration curve. hDiffusion-ordered
NMR spectroscopy (DOSY) of the alpha1 peptide, alpha1oleate complex, human serum albumin (HSA), and oleate suspension.
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
6NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
gap junction protein that has been proposed to promote cancer
development and metastasis33,34. The effect occurred specically
in tumor tissue, potentially providing the alpha1oleate complex
with a mechanism to trigger cell shedding, as observed here. It is
interesting to speculate that cell shedding may serve as a tip-of
the icebergmarker of the profound changes in tumor biology
that include activation of programmed cell death, transcriptional
reprogramming, and inhibition of tumor progression.
Fig. 3 Free energy surface analyses of the peptide- and peptideoleate system. a,bSuperimposition of dihedral principal component analysis
(PCA) plots of alpha1 (black points) and alpha1oleate (red points) systems (a), and sar1alpha (cyan points) and sar1alphaoleate (magenta points)
systems (b). Principal component (PC)1 and PC2 represent the axes of the two greatest variances after mathematical transformation for dimension
reduction. cfFree-energy surfaces as a function of the rst two principal components for alpha1oleate (c), naked alpha1 (d), sar1alphaoleate (e), and
naked sar1alpha (f). The representative structures of peptides or peptideoleate complexes, along with their respective local minima annotations, are
colored from the N termini (blue) to the C termini (orange/brown). The free-energy surface of the alpha1oleate complex contains 2 minima basins, A1 and
B1, with A1 representing the major conformational ensemble. The free-energy surface of the sar1alphaoleate complex contains 3 minima basins, A3, B3,
and C3 (with the A3 basin harboring the major structural ensemble), which are characterized by a prominent alpha-helical secondary structural element, as
shown from simulation calculated alpha-helical propensities. By contrast, the free-energy surface of the naked sar1alpha shows large structural
heterogeneity. Here, minima basins A4 and D4 are represented by helical structures, B4 by beta structure, and C4 and E4 by random coil structures.
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Fig. 4 Clinical study protocol, demographic data, and adverse events. a Study CONSORT diagram.bStudy protocol. After diagnosis and informed
consent, the subjects received intravesical instillations of either alpha1oleate or placebo on six occasions during one month preceding a scheduled
transurethral resection (TURB). A safety follow-up was performed 52 days after the rst instillation. cNumber of adverse events (AEs) in the active and
placebo groups. No drug-related adverse events were recorded. There were totally 29 AEs reported by 12 subjects in the active group and by 11 subjects in
the placebo group. None of the AEs were related to the investigational product. One AE was severe and two were moderate in the placebo group. The
active group had one moderate AE. Two subjects in the placebo group reported severe AE (SAEs). The AEs were evaluated descriptively, and the AE
proles were similar between the placebo and the active groups.
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
8NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Alternative therapeutic tools are actively being developed and
tested in patients with NMIBC, particularly in patients with
disease recurrence after BCG treatment3537. Device-assisted
hyperthermia was shown to increase the efcacy of intra-vesical
chemotherapy but treatment was accompanied by side effects,
reducing compliance3840. An oncolytic-virus-based intra-vesical
therapy was recently reported to achieve a complete response in
53.4% of patients with BCG-unresponsive carcinoma in situ, in a
phase III trial41. The authors discuss the assessment of side
effects and the development of biomarkers to help select patients
suitable for this therapy. In patients with BCG unresponsive
disease, treated with systemic Pembrolizumab, a 41% response
rate was reported but side effects were prevalent, limiting com-
pliance (Keynote-676 trial35). The present study identies
alpha1oleate as an active drug candidate with low toxicity.
Further dose-nding clinical studies and adjuvant therapy pro-
tocols will be essential to dene the therapeutic window of this
complex.
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Cancer cells are aggressive, outcompete healthy cells, and ruin
tissue integrity. It is generally assumed that treatments must be
equally aggressive and highly toxic substances are often used,
despite their lack of selectivity and the severe side effects that they
cause. The proteinlipid complexes studied here are attractive to
cancer cells, which actively internalized them, but end up being
killed. Healthy cells are less responsive and extensive toxicity
studies have failed to detect adverse effects in the bladder13. This
low toxicity was conrmed here, as no drug-related side effects
were observed in the treatment group. The results, therefore,
identify the alpha1oleate treatment of NMIBC as an interesting
therapeutic option. In view of the low toxicity observed so far,
liberal intra-vesical administration in early-stage NMIBC might
be an interesting approach to postponing the introduction of
more toxic and invasive therapeutic options.
Methods
Clinical trial of alpha1oleate in NMIBC
Trial design and patient population. This was a single-center, placebo-controlled,
double-blinded randomized Phase I/II interventional clinical trial of non-muscle
invasive bladder cancer taking place from May 21, 2018 to June 3, 2019. Subjects
diagnosed with non-muscle invasive bladder cancer and scheduled for transure-
thral surgery were included in the study. The study was approved by the State
Institute for Drug Control (SUKL) in the Czech Republic; number 273799/17-I and
the Ethics Committee of the Motol University Hospital; number EK-786/17
(ClinicalTrials.gov Identier: NCT03560479). Patients gave their written informed
consent.
Peptide synthesis and alpha1oleate complex generation. Peptide synthesis and the
preparation of the investigational product were performed under good manu-
facturing practice (GMP) conditions and the complex was diluted in phosphate-
buffered saline (PBS) to the nal concentration (1.7 mM). The placebo group
received PBS (sodium chloride, potassium chloride, sodium- and potassium
phosphate, and water for injection), which was identical in appearance to the active
treatment.
Study protocol. Study subjects were randomized 1/1 and received intra-vesical
instillations (30 mL) of either alpha1oleate (1.7 mM) or PBS on six occasions
during a period of 22 days (days 1, 3, 5, 8, 15, and 22). A safety follow-up was
included 52 days after the last instillation (EudraCT Number: 2016-004269-14 and
ClinicalTrials.gov NCT03560479). The complete Study Protocol is provided as
Supplementary Note 1 in the Supplementary Information le. The interim analysis
of the clinical trial presented here represents a complete evaluation of the Phase I/II
study, as per the original protocol. The protocol was later amended to include a 24-
month follow-up. At that time, it was decided to perform the initially planned nal
analysis as an interim analysis, scheduled after all subjects had completed the 52-
day safety follow-up. The study underwent data lock and subsequent unblinding
was under third-party control. The primary objective of the trial was to evaluate the
safety of alpha1oleate. No formal sample size calculation evaluating the power of
the trial has been performed. However, consideration regarding the sample size was
made based on a previous open study of HAMLET instillations in bladder cancer
patients16 and in the murine bladder cancer model14. For efcacy, the sample size
was based on analysis of change in tumor cells assessed before HAMLET
instillation and after 2 h. The mean fold increase of shed cells was 41.3 and the
standard deviation was 60.4 in 9 patients16. A sample size of 20 patients per group
was deemed suitable to achieve criterion for signicance (alpha) 0.05 and power
90% using the paired samples 1-tailed t-test. The null hypothesis is H
0
: mean
change in cell shedding =0 and the alternative hypothesis is H
A
: mean change in
cell shedding > 0.
Demographic data, morbidity and health parameters as well as tumor
characteristics were recorded by the study physicians in the electronic Case Report
Form (eCRF) and closely monitored by an external monitor. Population
characteristics were evaluated by the study statistician. No signicant differences
between the treatment and placebo groups were registered in terms of age, gender,
co-morbidity, or tumor parameters.
Primary endpoints. - Safety as AEs prole (time frame: from the signing of
informed consent (day 1) and until end of study (day 52)): Incidence of AEs and
classication in terms of severity, causality, and outcome.
-Efcacy as cell shedding (time frame: days 122): change in cell shedding into
urine (number of epithelial cells per mL of urine).
- Change from baseline in characteristics of papillary tumors (time Frame: prior
to treatment (baseline) and on day 30, in connection with scheduled surgery): the
bladder tumors are characterized by in vivo imaging during examination by
cystoscopy.
Secondary endpoints. - Histopathology scoring of the tumor using established
parameters for scoring of Grade and Stage/Invasiveness.
- Urine cytology examined before and after instillation, using the Paris scoring
system.
- Uptake of alpha1oleate by tumor cells, dened by staining with specic
antibodies.
- Tissue apoptotic response to alpha1oleate, dened by TUNEL staining.
- Tumor response to alpha1oleate, dened by RNA sequence analysis.
- Proteomic analysis of markers in urine was not completed.
- Long-term effects of the study treatment have not been evaluated.
AEs prole. AEs were collected from the signing of the informed consent form until
the end of the study (FU1 Visit, day 52). All diagnoses, symptom(s), sign(s), or
nding(s) with a start date after the rst dose of the study drug were recorded as
AEs or severe AEs (SAEs). (S)AEs related to the study procedure were coded
during the course of the trial according to MedDRA by preferred terms and pri-
mary system organ class. All AEs recorded during the course of the trial were
included in the subject data listings and an overall summary of the number
(percentage) of subjects with any treatment-emergent (S)AEs, premature dis-
continuations from the trial due to AEs, treatment-related AEs, and SAEs were
constructed. The number of subjects experiencing each type of adverse event was
tabulated regardless of the number of times each adverse event was reported by
each subject. The severity of each type of adverse event was also tabulated and
graded as the most severe recording for that adverse event.
Cell shedding. To quantify the shedding of cells and cell clusters into the urine,
samples were obtained from each patient prior to and after each instillation of
alpha1oleate or placebo (Visit 1Visit 6). Cell shedding was quantied by
counting the total number of epithelial cells in a unit of uncentrifuged urine under
light microscopy, using a hemocytometer chamber. Changes in cell shedding were
quantied at each visit, by comparing cell numbers in samples obtained before and
after each instillation. The cell clusters were scored based on the examination of
these samples by an experienced pathologist on a range of 02 where 0 =no
clusters and 2 =the highest number of clusters.
Fig. 5 Primary endpoints: shedding of tumor cells and reduction in tumor size following intra-vesical instillation of alpha1oleate. acCell shedding
increased signicantly after alpha1oleate instillation. aScatterplot showing individual means of six visits per patient in the treatment group (n=20)
compared to patients receiving placebo (n=20). Line represents the median. bComparison of cell numbers in urine before (pre =white) and after (post
=black) alpha1oleate inoculation on visits 16 showing increased cell numbers post-inoculation in the treatment group (n=20 patients per group, P=
0.0030 for visit 1, 0.0098 for visit 2, <0.0001 for visits 3 and 4, 0.0073 for visit 5 and 0.0336 for visit 6) but not in the placebo group. Data are presented
as mean ± SEM. cRepresentative images, illustrating the increase in cell shedding after alpha1oleate instillation. Magnication =×400. Scale bar =50 μm.
dfDifference in the shedding of tumor cell clusters between the treatment and placebo group. dScatterplot showing individual means of six visits per
patient in the treatment group compared to patients receiving placebo. Line represents the median. eIncreased numbers of cell clusters in post-inoculation
samples of patients receiving alpha1oleate (n=20 patients per group, P=0.9743 for visit 1, 0.0212 for visit 2, <0.0001 for visits 3, 4, and 5, and 0.0005
for visit 6). Data are presented as mean ± SEM. fRepresentative images of cancer cell clusters after alpha1oleate instillation. Magnication =×400. Scale
bar =50 μm. gParis grade of shed cells before or after alpha1oleate instillation. An increase is observed in the treatment group (χ2test). hReduction in
tumor size in patients receiving alpha1oleate treatment. Images were compared between the time of diagnosis and the time of TURB (P=0.04, χ2test for
trend compared to placebo, n=19 for treatment group and n=20 for placebo group). iExamples of cystoscopy photographs obtained by A.B. at the time
of diagnosis and after treatment at the time of TURB. Scale bars =5 mm. *P< 0.05, **P< 0.01, ***P< 0.001, ****P< 0.0001. The data were analyzed by
two-tailed unpaired MannWhitney U-test (a,d) or by repeated-measures two-way ANOVA with Sidaks correction (b,e).
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
10 NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Characteristics of papillary tumors, tumor size. To examine if the alpha1oleate
treatment affects tumor size, all included subjects underwent outpatient cystoscopy
at Visit 0. Tumors were reexamined at Visit 7, prior to scheduled surgery. High-
quality photographs were collected endoscopically, using a exible cystoscope
(Olympus) before being removed by TURB according to EAU Guideline
recommendations42. Changes in tumor size were evaluated intra-individually,
using paired images. The results were evaluated using a simplied Delphi
method23, by an independent NMIBC expert. Changes in lesion size, supercial
necrosis, and tissue vascularization were addressed in a blinded manner.
Histopathology scoring. Tumor biopsies, collected at the time of surgery were
evaluated by histopathology, using established parameters for scoring of Grade and
Stage/Invasiveness. Tissue samples were analyzed by a designated study uro-
pathologist. Both grading classications (WHO 1973 and 2004/2016) were used43.
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 11
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Biopsies from healthy tissue areas distant from the tumor were collected for
comparison.
Urine cytology. Urine cells were centrifuged onto L-lysine-coated microscope slides
(Cytospin 3, Shandon) at 113×gfor 5 min, xed and stored at room temperature
until further analyses. Urinary cytology was evaluated using the Paris System for
Reporting Urinary Cytology 201643,44 and dened as: 1. No diagnosis/unsatisfac-
tory. 2. Negative for high-grade urothelial carcinoma. 3. Atypical urothelial cells
present. 4. Suspicious for high-grade urothelial carcinoma. 5. High-grade urothelial
carcinoma. 6. Low-grade urothelial neoplasm. 7. Other positive for malignancies
and miscellaneous lesions.
Alpha1oleate uptake. Alpha1oleate uptake by tumor cells was quantied by
staining with specic antibodies. Cells on cytospin-slides were washed (Tris-buffer
saline TBS, 10 min), permeabilized (0.25% TritonX-100 in TBS, 20 min, room
temperature) and blocked (5% normal goat serum in TBS, 1 h, room temperature)
before the addition of rabbit polyclonal anti-human alpha-lactalbumin antibodies
(1:50 in 5% normal goat serum at 4 °C, overni ght, Mybiosource, Cat# MBS175270).
Slides were washed (TBS, 2 ×5 min) and stained with Alexa-568-labeled secondary
antibody (1:200, 1 h, room temperature, ThermoFisher). The nucleus was coun-
terstained using DRAQ5 (1:1000, 15 min) before a nal wash (2 ×5 min in TBS).
Slides were mounted (Fluoromount aqueous mounting media), before capturing
images by laser scanning confocal microscopy (Carl Zeiss). Fluorescence intensity
was quantied by ImageJ and net uorescence calculated after subtraction of the
secondary antibody background.
Apoptosis in tissue biopsies, TUNEL staining. DNA fragmentation was detected
using the terminal deoxynucleotidyl transferase dUTP nick end-labeling (TUNEL)
assay (Click-iT TUNEL Alexa Fluor 488 imaging assay kit, ThermoFisher). Tissue
sections were de-parafnized with xylene followed by serial dehydration with
ethanol (100%, 95%, 75%, and 50%). Dehydrated sections were xed (4% PFA, 15
min), permeabilized (DNase-free Protease K solution 20 µg/mL, 15 min) and
incubated with TUNEL reaction mixture containing TdT for 60 min at 37 °C. After
the TUNEL reaction, sections were incubated with Click-iT reaction mixture (30
min, 37 °C). Sections were counterstained with DAPI (1 µg/mL, 5min), mounted in
Fluoromount aqueous mounting media, and analyzed by uorescence microscopy
(Zeiss). Fluorescence intensities were quantied by ImageJ and net mean uores-
cence intensity calculated after subtraction of background uorescence.
RNA sequence analysis of tissue biopsies. RNA was extracted from tissues
stabilized in RNAlater using the AllPrep DNA/RNA/miRNA Universal Kit. Dis-
ruption was in the TissueLyser system and CK28 Precellys tubes and by homo-
genization in the QIAshredder homogenizer. The quantity and quality of the RNA
samples were evaluated using NanoDrop and Agilent 2100 Bioanalyzer. RNA
samples were prepared by Illumina TruSeq Stranded mRNA Library Prep Kit
(20020594), and libraries were multiplexed and sequenced using NextSeq 500/550
High Output Kits (v2.5 2 × 75 Cycles) with an average of 22 million reads per
sample. Raw sequencing data were demultiplexed using bcl2fastq (version 2.18)
and RSEM (1.3) was used for abundance estimation using the human genome
release 37/Ensemble 75. Samples were thoroughly quality checked (QC) and
visualized using dimensionality reduction (i.e. PCA), MA-plots as well as RNA-seq
intrinsic biases (such as GC bias, transcriptome complexity, and alignment quality).
Differential expression analysis was performed using R (version 3.4) and the
packages limma and DESeq2. Fold changes were calculated by comparing tumors
in the treated to the placebo group. Relative expression levels were analyzed and
genes with an absolute fold change >2.0 and P< 0.05 were considered as differ-
entially expressed. Heat-maps were constructed using the Gitools 2.1.1 software.
Differentially expressed genes were functionally characterized using the Ingenuity
Pathway Analysis version 57662101 (IPA, Qiagen) software.
Statistical analysis. For efcacy, the sample size was based on analysis of change
in tumor cells assessed from a previous study16. A sample size of 20 patients per
group was deemed suitable to achieve criterion for signicance (alpha) 0.05 and
power 90% using the paired samples 1-tailed t-test. The null hypothesis is H
0
: mean
change in cell shedding =0 and the alternative hypothesis is H
A
: mean change in
cell shedding > 0. The Gaussian distribution was determined by the Dagostino and
Pearson normality test. For data following a Gaussian distribution, student t-tests
were used. Other data sets were analyzed by MannWhitney U-test. Correlations
were determined by Spearman correlation. Kinetic data were analyzed using the
repeated measures two-way ANOVA test. All statistical analysis was done by using
Prism version 6.02 (GraphPad Software Inc.). Pvalues < 0.05 were considered
statistically signicant. All images were created by the study team.
Molecular and cellular studies
Chemicals and antibodies. Sodium oleate (Sigma-Aldrich, Cat# O7501), Alexa-
Fluor568 protein labeling kit (Thermo Scientic, Cat# A10238), AlexaFluor488
protein labeling kit (Thermo Scientic, Cat# A10235), ATPlite (Perkin Elmer, Cat#
6016947), Presto Blue Cell Viability Assay (Invitrogen, Cat# A13262), Anti-peptide
antibodies (this study, produced by GeneCust), FluxOR Potassium ion channel
assay (Invitrogen, Cat# F20015), Barium chloride BaCl
2
(Sigma-Aldrich, Cat#
B0750), Amiloride (Sigma-Aldrich, Cat# A7410), Click-iT TUNEL Alexa Fluor 488
imaging assay kit (ThermoFisher Scientic Cat# C10245), DRAQ5 (Abcam, Cat#
ab108410), Fluoromount (Sigma-Aldrich, Cat# F4680), DNA/RNA/miRNA Uni-
versal Kit (Qiagen, Cat# 80224),
Peptide synthesis and complex generation. Peptides for in vitro and in vivo
experiments were synthesized using Fmoc solid-phase chemistry (Mimotopes,
Melbourne, Australia). For biotinylated peptides, an aminohexanoic acid (Ahx)
spacer was added to ensure adequate separation between the biotin and the peptide
moieties. A ve-fold stoichiometric concentration of sodium oleate in phosphate-
buffered saline was prepared, followed by the addition of each respective peptide.
The more hydrophobic peptides were initially dissolved in DMSO, then transferred
to the oleate buffer. The sequences for the peptides are as follows:
Alpha1: Ac-KQFTKAELSQLLKDIDGYGGIALPELIATMFHTSGYDTQ-OH
Beta: Ac-IVENNESTEYGLFQISNKLWAKSSQVPQSRNIADISADKFLD
DD-OH
Sar1alpha: Ac-MAGWDIFGWFRDVLASLGLWNKH-OH
Sar1beta: Ac-DRLATLQPTWHPTSEELAIGNIKFTTFDLGGHI-OH
Cell lines and cell culture. Human lung carcinoma cells (A549, ATCC Cat# CCL-
185, RRID:CVCL_0023), human kidney carcinoma cells (A498, ATCC Cat# HTB-
44, RRID:CVCL_1056), and murine bladder carcinoma cells (MB49, RRID:
CVCL_7076, provided by Sara Mangsbo, Uppsala University, Sweden) were cul-
tured in RPMI-1640 with non-essential amino acids (1:100), 1 mM sodium pyr-
uvate, 50 μg/mL gentamicin, and 510% fetal calf serum (FCS) at 37 °C, 5% CO
2
.
Cell death assays. To quantify effects on cell viability, A549, A498, or MB49 cells
were seeded in 96-well plates (2 × 104/well, Tecan Group Ltd.), cultured overnight
at 37 °C, 5% CO
2
and incubated with peptideoleate complexes in serum-free
Fig. 6 Apoptotic response to alpha1oleate and cellular uptake by tumor cells. Apoptosis was quantied in tumor biopsies, using the TUNEL assay.
Arbitrary units were calculated after subtraction of background staining in TUNEL negative healthy tissue samples. aRepresentative image of TUNEL
staining (green =TUNEL, blue =DAPI) in tumor tissue from individual patients receiving alpha1oleate instillations. Scale bars =200 μm. bRepresentative
images of TUNEL staining in tumor tissue from individual patients receiving placebo. Scale bars =200 μm. cScatter plot demonstrating elevated TUNEL
staining intensity in tumor biopsies from patients receiving alpha1oleate instillations compared to placebo. TUNEL staining was not signicantly altered in
healthy tissue biopsies from patients receiving alpha1oleate instillations or placebo (n=40 tumors and 38 healthy biopsies, two data points were further
removed due to medical conditions from patients and conrmed by Grubbss outlier test) (two-tailed unpaired MannWhitney U-test). Line represents the
median. dCorrelation of TUNEL staining intensity with cell shedding (P=0.03, 95% CI 0.02200.6010) and alpha1oleate uptake (P=0.01, 95% CI
0.09570.6461) (Spearman correlation, two-tailed, approximate P-value, n=20 for alpha1o and n=19 for placebo). eRepresentative images of
alpha1oleate (red) uptake with counter-stained nuclei (blue). Alpha1oleate uptake by tumor cells was quantied by staining of shed cells in urine with
polyclonal anti-alpha1oleate antibodies. Scale bars =20 μm. fScatterplots of cellular uptake in individual patients receiving alpha1oleate. Each dot
represents the mean uorescence intensity of six post-instillation samples per patient treated with alpha1oleate. Comparison of alpha1oleate uptake by
the cell in urine before (pre =white) and after (post =black) alpha1oleate inoculation on visits 1-6 (repeated-measures two-way ANOVA with Sidaks
correction, P=0.0049 for visit 1, 0.1913 for visit 2, 0.0067 for visit 3, 0.0025 for visit 4, 0.3807 for visit 5 and 0.0043 for visit 6, n=20 per group and
time point). Line represents the median and bars represent mean ± SEM.
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
12 NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
RPMI-1640 at 37 °C. FCS was added after 1h and cell viability was quantied after
3 h, by the Presto Blue uorescence assay (Thermo Scientic) or ATP levels
(luminescence-based ATPlite kit), using a microplate reader (Innite F200, Tecan).
The colony assay was used to dene long-term effects on cell viability. MB49
cells were seeded in a 24-well tissue culture plate (1000 cells/well in 1 mL in RPMI
with FCS), allowed to adhere for 24 h, was hed in PBS and treated with
alpha1oleate or sar1alphaoleate (7, 21, or 35 μM in 1 mL of RPMI without FCS
for 1 h). After the addition of FCS (5%) the cells were incubated at 37°C in 5% CO
2
for 10 days or until visible colonies were present in the PBS control. Plates were
xed in cold methanol, stained with hematoxylin. Staining intensity was quantied
using ImageJ.
TUNEL staining was performed using Click-iT TUNEL Alexa Fluor 488
Imaging Assay kit (Thermo Scientic). Briey, the cells were xed in 2%
paraformaldehyde for 15 min followed by permeabilization with 0.25% TritonX-
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 13
Content courtesy of Springer Nature, terms of use apply. Rights reserved
100 for 20 min. The sections were incubated for 60 min at 37 °C with the TUNEL
reaction mixture, then counter-sta ined with DAPI (20 μg/mL) for 15 min. A
positive contro l slide (cells treated with DNase I for 30 min at 37 °C) was included
in each experiment. The TUNEL uorescent intensity was quantied using Image J.
Membrane blebbing and ion uxes. Membrane changes in lung carcinoma cells
were visualized by transmission light microscopy imaging. Cells were seeded on a
glass coverslip and allowed to partially adhere to the glass surface for 10 min at
room temperature, prior to exposure to the alpha1oleate or sar1alphaoleate
complexes. Changes in cell morphology were captured with LSM 510 META
confocal microscope (Carl Zeiss) using ×40 oil immersion objective. Naked pep-
tides and oleate were used as negative controls (Supplementary Fig. 1e).
K+uxes were quantied using the FluxOR potassium ion channel assay
(Invitrogen). The inux of the indicator Ti+measures the opening of K+
channels45. Briey, 20,000 adherent cells were incubated in loading buffer followed
by incubation with assay buffer and stimulus buffer for 60 min. For inhibition, cells
were pretreated with the K+channel inhibitor BaCl
2
(100 μM) followed by
treatment with alpha1oleate (35 μM). Fluorescence was measured at 535 nm after
excitation at 485 nm using a uorescence plate reader (TECAN innite F200).
Structural analysis of the peptideoleate complexes
CD spectroscopy. Far-ultraviolet (UV) CD spectra were collected on alpha1-, beta-,
sar1alpha-, and sar1beta-peptides with and without oleate at 25 °C using a Jasco
815 CD Spectropolarimeter. The peptides were dissolved in 50 mM sodium
phosphate buffer, pH 7.4, with 10% D
2
O, at a nal concentration of 0.2 mg/mL.
Far-UV CD was performed from 185 to 260 nm for the samples without oleate and
from 200 to 260 nm for the samples with oleate and the background was sub-
tracted. The mean residue ellipticity (MRE), [θ], in deg cm2/dmol, was calculated as
described previously9.
Biomolecular NMR spectroscopy. The alpha1- and sar1alpha-naked peptide samples
were dissolved in 50 mM sodium phospha te buffer (pH 7.4, 90% H
2
O, 10% D
2
O),
and the peptideoleate complexes were reconstituted from a lyophilisate of
phosphate-buffered saline. All experiments were carried out in the phase-sensitive
mode46. One-dimensional 1H, two-dimensional NOESY (Nuclear Overhauser
Effect Spectroscopy) and 1H13C HSQC (Heteronuclear Single Quantum Corre-
lation) spectra were acquired on an Agilent Technologies 18.8T (800 MHz) DD2
Premium Compact spectrometer with a triple-resonance, 5 mm enhanced cold
probe. The 1H13C HSQC spectra were collected at 20 °C with 16 scans, an initial
delay of 3.0 s, a 90° pulse width of 7.5 and 9.8 μs, and an acquisition time of 0.4 s
with broadband decoupling for alpha1 and sar1alpha peptide samples. For the
alpha1oleate and sar1alphaoleate complexes, we used a 90° pulse width of 12.80
and 13.30 μs and an acquisition time of 0.4 s. All acquisition parameters were kept
constant for all samples. Two-dimensional DPFGSE-NOESY (Double Pulse Field
Gradient Spin Echo-NOESY) pulse sequences were used to acquire data at 20 °C
with 16 scans, with an optimized mixing time of 300 ms for the alpha1 and
alpha1oleate complexes and a delay period of 1.5 s. For the sar1alphaoleate
complex, water-gate NOESY was used with 12 scans, with a mixing time of 150 ms.
A trace amount of TSP was added to serve as a chemical shift reference. Each 2D
HSQC spectrum consisted of 4 K complex points in the acquisition dimension and
512 complex points in the indirect dimension. For the NOESY spectra, 4 K com-
plex points were used in the acquisition dimension and 1 K complex points in the
indirect dimension. The two-dimensional data were processed with Gaussian
apodization in both dimensions. The stoichiometry of the peptide with the oleic
acid was determined by comparing the peak areas (using the 1D 1H spectra) or
peak volumes (using the 2D 1H13C HSQC spectra) of well-resolved, isolated
regions found in the spectra.
Diffusion-ordered spectroscopy (DOSY) measurements were performed at 293
K. Samples were prepared in 50 mM phosphate buffer at pH 7.4. The
DgscteSL_dpfgsc DOSY pulse program was used, which consists of gradient
compensated stimulated echo with spin lock using the excitation sculpting solvent
suppression method47. A spectral window of 13,020 Hz was used, with an
acquisition time of 2.46 s with a relaxation delay of 3 s. The FIDs were collected
with 32,000 complex data points with 64 scans. Logarithmically the gradient pulse
strength was increased from 3% to 86% of the maximum strength of 32,767 G/cm
in 60 steps. A diffusion time (Δ) of 100 ms and bipolar half-sine-shaped gradient
pulses (δ) of 5 ms was applied. 1,4-Dioxane, which is known to behave
independently of protein concentration and the folded state of the protein, was
used as an internal chemical shift reference and hydrodynamic radius calibration
reference (3.75 ppm; R
H
=2.12 Å)48,49. DOSY processing was performed using a
two-component t with a discrete approach, which further processed using a non-
uniform gradients approach. Three replicate acquisitions were given for each
sample, and the resulting diffusion coefcient (D) values calculated. For alpha1
peptide the average Dvalue was 2.162 and 14.10 m2/s for 1,4-dioxane. In the case
of alpha1oleate complex the average Dvalue was 0.986 m2/s for complex and
13.61 m2/s for 1.4-dioxane. The calculated R
H
are as follows: alpha1 peptide R
H
=
13.82 ± 0.447 Å, alpha1oleate complex R
H
=29.3 ± 0.606 Å, HSA R
H
=40.9 ±
1.44 Å, oleate in aqueous solution R
H
=104.3 ± 7.22 Å, oleate in methan ol R
H
=
5.58 ± 0.0649 Å. Note that the D(diffusion coefcient) values for 1,4-dioxane are
slightly variable dependent upon the co-solute (lower panel where Dis between
14.4 and 12.6), which rightly reects the different solution micro-environment
conditions that both solutes are mutually experiencing for each sample.
For T
2
relaxation measurements, the standard CPMGT2 pulse sequence was
used to run the experiments with 15 relaxations delays, which were chosen
logarithmically for different maximum T
2
time intervals: 8 s (alpha1 peptide), 1.2s
(alpha1oleate complex), 3.0 s (HSA), 7.0 s (oleate in aqueous solution), and 10 s
(oleate in methanol), respectively. The data were acquired with 32,000 complex
points with a baseline correction of 4. The T
2
analyses were performed on VNMRJ
version 4.0 (Agilent Technologies) software by the exponential tting of these
values with their corresponding intensity. All other NMR parameters were kept
constant for all samples throughout the experiments. The experiments were
acquired at a sample temperature of 293 K. The data are presented in
Supplementary Table 1.
Size-exclusion HPLC. Calibration standards and samples were injected onto a
TSKgel Super SW3000 HPLC column (4.6 mm × 30 cm, Particle size 4 μm, pore
size 25 nm, Tosoh Bioscience) eluted with 0.05 M sodium phosphate buffer pH 7.0
containing 0.1 M Na
2
SO
4
at a ow rate of 0.25 mL/min and detection at 280 nm.
The chromatography was performed on a Dionex Ultimate HPLC 3000 Standard
System running Chromeleon 6 software (Dionex, Thermo Scientic). The standard
calibration curve was generated with the proteins given including HSA, which has a
hydrodynamic radius (R
H
) of 40 Å (Supplementary Fig. 5). The R
H
vs. elution
volume linearity of the standard calibration curve is known to vanish after
approximately 3.8 mL of elution volume50. As a result, the R
H
of oleate in methanol
eluent (a solvent that ensures that the fatty acid is monomeric) will be less than
what is estimated from the standard curve (12.3 Å). The retention times of small
R
H
-analytes are closely reproduced regardless of eluent, be it aqueous buffer or
methanol.
Computational simulations: model building of peptide and peptideOA complexes.
The initial structure of the alpha1 peptide was obtained from the corresponding
domain in the crystal structure of human alpha-lactalbumin (PDB ID: 1B9O). All
cysteines were mutated to alanines, consistent with ndings that a reduced human
alpha-lactalbumin mutant in which all cysteines mutated to alanines could form a
cytotoxic complex in the presence of the lipid cofactor9. The initial structure of the
sar1alpha peptide was obtained from an I-TASSER-built homology model51. The
sequence similarity of the Sar1alpha peptide with the sequence of the top-ranked
threading template used by I-Tasser (PDB ID: 1R7G) was 0.35. However, this is
eventually irrelevant as we used extensive H-REMD sampling to sample the con-
formations of the peptide, both in its oleate bound and apo forms. The alpha1 and
the sar1alpha peptide were centered in a cubic box with box edges 1.2 nm from the
peptide. For the oleate-containing systems of alpha1 and sar1alpha, 4 molecules of
oleate are placed randomly in the box surrounding the alpha1 peptide to obtain a
peptideoleic acid ratio of 1:4. The Amber 99SB-ildn52 force eld and the TIP3P53
water model were used. For the coordinates, the starting structure was built using
Discovery Studio 4.1 (Accelrys). Geometry optimization for the ligand was per-
formed using Gaussian0954 at the level of HF-6-31G*, and the partial charges were
determined by the RESP55 method implemented in the antechamber tool of
AmberTools16 (AMBER 2016). Topologies for the oleate were built using the
General Amber Force Field56. The respective atom labels, corresponding atom type,
and partial charges are shown in Supplementary Table 10. Additional parameters
Fig. 7 Reprogramming of gene expression. RNA sequencing was used to compare gene expression proles in tumor tissue biopsies from the treatment or
placebo groups. aPie chart of genes regulated in response to treatment (cut-off FC > 1.5, P< 0.05 compared to the placebo group). In the treatment group,
82% of all regulated genes were cancer-related and 14% were bladder cancer-related. Gene categories were identied by biofunction analysis. bHeatmap
of specic cancer- and bladder cancer-related genes regulated in tumor biopsies from the treatment group (red =upregulated, blue =downregulated, cut-
off FC > 1.5, P<0.05 compared to placebo group). About 60% of all regulated genes were inhibited in the treatment group. cDetailed analysis of data in (a,
b). Top regulated, cancer-associated functions are shown. Inhibition is indicated by negative z-scores (blue) and signicance by Pvalues (orange). The
expression of genes involved in tumor invasion, neoplasia, tumor growth, and urinary tract tumors was strongly inhibited. dInhibition of Ras signaling in the
treatment group compared to placebo. eBladder cancer gene network regulated specically in patients receiving alpha1oleate treatment compared to
placebo.
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
14 NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
for the molecule are derived from the General Amber Force Field. For reference,
the structure of the oleate molecule and its atom labels is shown in Supplementary
Fig. 11. All systems were neutralized and Na+and Clions were added to a
concentration of 0.15 M. Energy minimization was performed using the steepest
descent algorithm for 1000 steps to remove any initial bad contacts. Long-range
electrostatics were treated with the particle mesh Ewald algorithm57, with a real-
space cutoff of 1.2 nm, and Van der Waals interactions were truncated at 1.2 nm.
All systems with oleate-containing peptides or naked peptides were initially heated
at 500 K for 40 ns to eliminate starting structure bias and provide a partially
unfolded state for the peptides. Temperature coupling of the system was performed
using a velocity rescaling thermostat58.
Hamiltonian replica exchange molecular dynamics simulations. The Gromacs 5.1.2
molecular dynamics package56 with the Plumed 2.3 plugin for Hamiltonian Replica
Exchange Molecular Dynamics59 was used to perform the simulations. All atoms of
the alpha1- and sar1alpha peptide residues, along with the oleate residues of the
two oleate-containing systems, were selected for Hamiltonian scaling. Twenty
replicas were used for each system, and scaling factors were generated for an
effective temperature range of 300800 K. Temperatures for scaling were selected
based on a geometric progression. The temperature factors were 300, 315.893,
332.629, 350.251, 368.807, 388.346, 408.919, 430.583, 453.395, 477.415, 502.707,
529.34, 557.384, 586.913, 618.006, 650.747, 685.223, 721.525, 759.75, and 800 K.
Each replica was simulated for 400 ns, resulting in an effective simulation of 8 µs.
Exchanges were attempted every 2 ps, and the result was an average acceptance
probability of approximately 30%.
Simulation analysis. Analysis of simulation data was performed using the built-in
Gromacs tools of the Gromacs package56. The ensemble for each system with the
canonical unscaled potential energy was used for the analysis, and data analysis was
performed on the last 300 ns for each system. Dihedral principal component
analysis60 was performed using the gmx angles, gmx covar, and gmx anaeig tools to
prepare and diagonalize the covariance matrix and analyze eigenvectors and
eigenvalues. The free-energy surface was constructed through projection onto the
rst and second principal components with the formula F
i
=RT ln(P
i
/P
o
), where
Ris the gas constant, Tis temperature (300 K), P
i
is the population in each bin and
P
o
is the population of the most populated bin. The gmx cluster tool of Gromacs
5.1.2 was used to identify the representative structure of each minima for geometric
clustering for the Gromos algorithm. We used the dene secondary structure of
proteins (DSSP)61 algorithm to calculate secondary structure propensities. For our
analysis, we classied the 3
10
helix, αhelix, and πhelix structures as helices; the β-
sheet and residue in isolated β-bridge structures as sheets and the remaining
structures as others. The contact probability was calculated using the gmx mindist
tool in the Gromacs package. The minimum distance between protons of side
chains for each residue and oleic acid was calculated for each frame. To calculate
the contact probability, a contact was dened if the measured distance was less than
0.55 nm. The contact probabilities between Aromatic ring protons and Olenic
protons of alpha1- and sar1alphaoleate-containing systems were also calculated
similarly. Proton distances were calculated to facilitate the comparison of simu-
lation data to Nuclear Overhauser Spectroscopy data.
Reporting summary. Further information on research design is available in the Nature
Research Reporting Summary linked to this article.
Data availability
The data supporting the structural and cellular ndings of this study are available within
the article and its supplementary information les. The structural data referenced during
the study are available in a public repository from the Protein Data Bank website (www.
rcsb.org, DOI:10.2210/pdb1B9O/pdb, DOI:10.2210/pdb1R7G/pdb). The RNA
sequencing data generated in this study have been deposited in the Gene Expression
Omnibus (GEO) database under accession number GSE172112. Source data are provided
with this paper.
Received: 8 April 2020; Accepted: 15 April 2021;
References
1. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell
144, 646674 (2011).
2. Scott, A. M., Wolchok, J. D. & Old, L. J. Antibody therapy of cancer. Nat. Rev.
Cancer 12, 278287 (2012).
3. Håkansson, A., Zhivotovsky, B., Orrenius, S., Sabharwal, H. & Svanborg, C.
Apoptosis induced by a human milk protein. Proc. Natl Acad. Sci. USA 92,
80648068 (1995).
4. Svensson, M., Håkansson, A., Mossberg, A.-K., Linse, S. & Svanborg, C.
Conversion of α-lactalbumin to a protein inducing apoptosis. Proc. Natl Acad.
Sci. USA 97, 42214226 (2000).
5. Gustafsson, L., Leijonhufvud, I., Aronsson, A., Mossberg, A. K. & Svanborg, C.
Treatment of skin papillomas with topical alpha-lactalbuminoleic acid. N.
Engl. J. Med. 350, 26632672 (2004).
6. Chiti, F. & Dobson, C. M. Protein misfolding, amyloid formation, and human
disease: a summary of progress over the last decade. Annu. Rev. Biochem. 86,
2768 (2017).
7. Knowles, T. P., Vendruscolo, M. & Dobson, C. M. The amyloid state and its
association with protein misfolding diseases. Nat. Rev. Mol. Cell Biol. 15,
384396 (2014).
8. Ho, C. S. J., Rydstrom, A., Manimekalai, M. S. S., Svanborg, C. & Grüber, G.
Low resolution solution structure of HAMLET and the importance of its
alpha-domains in tumoricidal activity. PLoS ONE 7, e53051 (2012).
9. Pettersson-Kastberg, J. et al. Alpha-lactalbumin, engineered to be nonnative
and inactive, kills tumor cells when in complex with oleic acid: a new
biological function resulting from partial unfolding. J. Mol. Biol. 394,
9941010 (2009).
10. Fjell, C. D., Hiss, J. A., Hancock, R. E. & Schneider, G. Designing antimicrobial
peptides: form follows function. Nat. Rev. Drug Discov. 11,3751 (2011).
11. Mok, K. H., Nagashima, T., Day, I. J., Hore, P. J. & Dobson, C. M. Multiple
subsets of side-chain packing in partially folded states of α-lactalbumins. Proc.
Natl Acad. Sci. USA 102, 88998904 (2005).
12. Fischer, W. et al. Human alpha-lactalbumin made lethal to tumor cells
(HAMLET) kills human glioblastoma cells in brain xenografts by an
apoptosis-like mechanism and prolongs survival. Cancer Res. 64, 21052112
(2004).
13. Hien, T. T. et al. Bladder cancer therapy without toxicitya dose-escalation
study of alpha1oleate. Int. J. Can. 147, 24792492 (2020).
14. Mossberg, A. K., Hou, Y., Svensson, M., Holmqvist, B. & Svanborg, C.
HAMLET treatment delays bladder cancer development. J. Urol. 183,
15901597 (2010).
15. Puthia, M., Storm, P., Nadeem, A., Hsiung, S. & Svanborg, C. Prevention and
treatment of colon cancer by peroral administration of HAMLET (human α-
lactalbumin made lethal to tumour cells). Gut 63, 131142 (2014).
16. Mossberg, A. K. et al. Bladder cancers respond to intravesical instillation of
(HAMLET human α-lactalbumin made lethal to tumor cells). Int. J. Can. 121,
13521359 (2007).
17. Barlowe, C. et al. COPII: a membrane coat formed by Sec proteins that drive
vesicle budding from the endoplasmic reticulum. Cell 77, 895907 (1994).
18. Jensen, D. & Schekman, R. COPII-mediated vesicle formation at a glance. J.
Cell. Sci. 124,14 (2010).
19. Sato, K. & Nakano, A. Dissection of COPII subunit-cargo assembly and
disassembly kinetics during Sar1p-GTP hydrolysis. Nat. Struct. Mol. Biol. 12,
167174 (2005).
20. Nadeem, A. et al. Protein receptor-independent plasma membrane
remodeling by HAMLET: a tumoricidal proteinlipid complex. Sci. Rep. 5,
16432 (2015).
21. Kang, M., Jeong, C. W., Kwak, C., Kim, H. H. & Ku, J. H. Single, immediate
postoperative instillation of chemotherapy in non-muscle invasive bladder
cancer: a systematic review and network meta-analysis of randomized clinical
trials using different drugs. Oncotarget 7, 4547945488 (2016).
22. Sylvester, R. J. et al. Systematic review and individual patient data meta-
analysis of randomized trials comparing a single immediate instillation of
chemotherapy after transurethral resection with transurethral resection alone
in patients with stage pTapT1 urothelial carcinoma of the bladder: which
patients benet from the instillation? Eur. Urol. 69, 231244 (2016).
23. Khare, S. R. et al. Quality indicators in the management of bladder cancer: a
modied Delphi study. Urol. Oncol. 35, 328334 (2017).
24. Ho, J., Nadeem, A., Rydström, A., Puthia, M. & Svanborg, C. Targeting of
nucleotide-binding proteins by HAMLETa conserved tumor cell death
mechanism. Oncogene 35, 897907 (2016).
25. Lawler, P. R. & Lawler, J. Molecular basis for the regulation of angiogenesis by
thrombospondin-1 and -2. Cold Spring Harb. Perspect. Med. 2, a006627
(2012).
26. Antoni, S. et al. Bladder cancer incidence and mortality: a global overview and
recent trends. Eur. Urol. 71,96108 (2017).
27. Sievert, K. D. et al. Economic aspects of bladder cancer: what are the benets
and costs? World J. Urol. 27, 295300 (2009).
28. van Rhijn, B. W. et al. Recurrence and progression of disease in nonmuscle-
invasive bladder cancer: from epidemiology to treatment strategy. Eur. Urol.
56, 430442 (2009).
29. Hu, Q., Sun, W., Wang, C. & Gu, Z. Recent advances of cocktail chemotherapy
by combination drug delivery systems. Adv. Drug Deliv. Rev. 98,1934 (2016).
30. Sharma, P. & Allison, J. P. Immune checkpoint targeting in cancer therapy:
toward combination strategies with curative potential. Cell 161, 205214
(2015).
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y ARTICLE
NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications 15
Content courtesy of Springer Nature, terms of use apply. Rights reserved
31. Ourfali, S. et al. Recurrence rate and cost consequence of the shortage of
Bacillus Calmette-Guérin connaught strain for bladder cancer patients. Eur.
Urol. Focus 7, 111116 (2021).
32. Dang, C. V. MYC on the path to cancer. Cell 149,2235 (2012).
33. Chevallier, D., Carette, D., Segretain, D., Gilleron, J. & Pointis, G. Connexin 43
a check-point component of cell proliferation implicated in a wide range of
human testis diseases. Cell. Mol. Life Sci. 70, 12071220 (2013).
34. Ghosh, S., Kumar, A. & Chandna, S. Connexin-43 downregulation in G2/M
phase enriched tumour cells causes extensive low-dose hyper-radiosensitivity
(HRS) associated with mitochondrial apoptotic events. Cancer Lett. 363,
4659 (2015).
35. Kamat, A. M. et al. KEYNOTE-676: Phase III study of BCG and
pembrolizumab for persistent/recurrent high-risk NMIBC. Future Oncol. 16,
507516 (2020).
36. Lamm, D. L. Intravesical therapy for supercial bladder cancer: slow but
steady progress. J. Clin. Oncol. 21, 42594260 (2003).
37. Kamat, A. M. et al. Evidence-based assessment of current and emerging
bladder-sparing therapies for nonmuscle-invasive bladder cancer after
Bacillus Calmette-Guerin therapy: a systematic review and meta-analysis. Eur.
Urol. Oncol. 3, 318340 (2020).
38. de Jong, J. J., Hendricksen, K., Rosier, M., Mostad, H. & Boormans, J. L.
Hyperthermic intravesical chemotherapy for BCG unresponsive non-muscle
invasive bladder cancer patients. Bladder Cancer 4, 395401 (2018).
39. Colombo, R. et al. Multicentric study comparing intravesical chemotherapy
alone and with local microwave hyperthermia for prophylaxis of recurrence of
supercial transitional cell carcinoma. J. Clin. Oncol. 21, 42704276 (2003).
40. Arends, T. J. H. et al. Results of a randomised controlled trial comparing
intravesical chemohyperthermia with mitomycin C versus Bacillus Calmette-
Guérin for adjuvant treatment of patients with intermediate- and high-risk
non-muscle-invasive bladder cancer. Eur. Urol. 69, 10461052 (2016).
41. Boorjian, S. A. et al. Intravesical nadofaragene radenovec gene therapy for
BCG-unresponsive non-muscle-invasive bladder cancer: a single-arm, open-
label, repeat-dose clinical trial. Lancet Oncol. 22, 107117 (2021).
42. Babjuk, M. et al. EAU guidelines on nonmuscle-invasive urothelial
carcinoma of the bladder: update 2016. Eur. Urol. 71, 447461 (2017).
43. Rosenthal, D. L., Wojcik, E. M. & Kurtycz, D. F. In The Paris System for Reporting
Urinary Cytology (eds Rosenthal, D. L., Wojcik, E. M. & Kurtycz, D. F.)
(Springer, 2016).
44. VandenBussche, C. A review of the paris system for reporting urinary
cytology. Cytopathology 27, 153156 (2016).
45. Storm, P. et al. A unifying mechanism for cancer cell death through Ion
channel activation by HAMLET. PLoS ONE 8, e58578 (2013).
46. States, D. J., Haberkorn, R. A. & Ruben, D. J. A two-dimensional nuclear
overhauser experiment with pure absorption phase in four quadrants. J. Magn.
Reson. 48, 286292 (1982).
47. Pelta, M. D., Barjat, H., Morris, G. A., Davis, A. L. & Hammond, S. J. Pulse
sequences for high-resolution diffusion-ordered spectroscopy (HR-DOSY).
Magn. Reson. Chem. 36, 706714 (1998).
48. Jones, J. A., Wilkins, D. K., Smith, L. J. & Dobson, C. M. Characterisation of
protein unfolding by NMR diffusion measurements. J. Biomol. NMR 10,
199203 (1997).
49. Shimizu, A., Ikeguchi, M. & Sugai, S. Appropriateness of DSS and TSP as
internal references for (1)H NMR studies of molten globule proteins in
aqueous media. J. Biomol. NMR 4, 859862 (1994).
50. SEC: Size Exclusion Chromatography (Tosoh Bioscience GmbH, 2017), https://
www.separations.eu.tosohbioscience.com/solutions/hplc-products/size-
exclusion.
51. Yang, J. et al. The I-TASSER Suite: protein structure and function prediction.
Nat. Methods 12,78 (2015).
52. Lindorff-Larsen, K. et al. Improved side-chain torsion potentials for the
Amber ff99SB protein force eld. Proteins 78, 19501958 (2010).
53. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M.
L. Comparison of simple potential functions for simulating liquid water. J.
Chem. Phys. 79, 926935 (1983).
54. Gaussian 09 (Gaussian, Inc., 2016).
55. Bayly, C. I., Cieplak, P., Cornell, W. & Kollman, P. A. A well-behaved
electrostatic potential based method using charge restraints for deriving
atomic charges: the RESP model. J. Phys. Chem. A 97, 1026910280 (1993).
56. Van Der Spoel, D. et al. GROMACS: fast, exible, and free. J. Comput. Chem.
26, 17011718 (2005).
57. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an Nlog(N) method
for Ewald sums in large systems. J. Chem. Phys. 98, 1008910092 (1993).
58. Bussi, G., Donadio, D. & Parrinello, M. Canonical sampling through velocity
rescaling. J. Chem. Phys. 126, 014101 (2007).
59. Wang, L., Friesner, R. A. & Berne, B. J. Replica exchange with solute scaling: a
more efcient version of replica exchange with solute tempering (REST2). J.
Phys. Chem. B 115, 94319438 (2011).
60. Mu, Y., Nguyen, P. H. & Stock, G. Energy landscape of a small peptide
revealed by dihedral angle principal component analysis. Proteins 58,4552
(2005).
61. Kabsch, W. & Sander, C. Dictionary of protein secondary structure: pattern
recognition of hydrogen-bonded and geometrical features. Biopolymers 22,
25772637 (1983).
Acknowledgements
The authors thank Petr Bouška and NEOX for the Monitoring of the clinical study;
Jeanette Valcich and the Center for Translational Genomics (CTG) at Lund University
for the mRNA Library prep and Sequencing; Susanne Strömblad, Lina Gefors and the
Lund University Bioimaging Centre (LBIC) for providing experimental resources for
tissue analysis, Arunima Chaudhuri for input and comments on the manuscript, Sara
Mangsbo, Uppsala University, Sweden for the MB49 cells (RRID:CVCL_7076). We
gratefully acknowledge the support of the Swedish Research Council, the Swedish Cancer
Society (Cancerfonden) and HAMLET Pharma. Support to the Svanborg group was
further provided from the European Unions Horizon 2020 research and innovation
program under grant agreement No. 954360. The funding sources had no role in the
design of this study or in its execution, analyses, interpretation of the data, or decision to
submit the results for publication.
Author contributions
All authors met the ICMJE criteria for authorship. C.S., M.B., J. Ho, Y.G.M., K.H.M.,
conceived and designed the study. J. Ho, P.S.K., A.H., D.L.F., K.H.M. performed struc-
tural biology experiments and analyses. J.T.Y.-N. performed and analyzed the
simulations under Y.G.M. supervision. J. Ho, A.N., T. Hiep T., P.E. performed
cellular experiments and analyses. A.B., J. Háček, I.A., D.S.C.B., M.L.Y.W., T. Hiep T.,
H.N., J. Horňák, M.B., and C.S. performed the human trial and analyses. M.B., A.B.,
J. Háček, I.A., D.S.C.B., T. Hiep T., T. Hien T., M.L.Y.W., P.S., M.B., C.S. analyzed the
data and J. Ho, I.A., D.S.C.B., T. Hiep T., K.H.M., M.L.Y.W., A.B., M.B., C.S. wrote
the paper.
Funding
Open access funding provided by Lund University.
Competing interests
C.S. holds shares in HAMLET Pharma, as a representative of scientists in the HAMLET
group. Patents protecting the use of the alpha1 peptide were led previously (Biologically
active complexes and therapeutic uses thereof; GB 201707715 priority date 14/05/2017,
PCT/EP2018/062396 ling date 14/05/2018; inventors: C.S., A.N., J.Ho). No specic
patents have been led based on this study. Other authors declare no competing or
conicts of interests.
Additional information
Supplementary information The online version contains supplementary material
available at https://doi.org/10.1038/s41467-021-23748-y.
Correspondence and requests for materials should be addressed to C.S.
Peer review information Nature Communications thanks Keith Syson Chan and the
other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer
reviewer reports are available.
Reprints and permission information is available at http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2021
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-021-23748-y
16 NATURE COMMUNICATIONS | (2021) 12:3427 | https://doi.org /10.1038/s41467-021-23748-y | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... By another hand, the diversity and the abundance of the gut microbiota imply a role in the human health [38] in the homeostasis of nutrients metabolism, and gut immunity [39;40]. The gut Figure 1A: The gut microbiota composition under normal conditions. ...
... Firm cutes are mainly Gram-positive bacteria, and mostly Clostridium (95%) and 5% are conformed by Lactobacillus, Bacillus, Eritrococcus and Ruminococcus. While the Bacteroidetes usually are Gram negative bacteria, such as Bacteroidetes and Prevotella [38]. Any disbalance onto the gut microbiota composition leads to disease in dysbiosis, which has been associated to inflammatory diseases such as, inflammatory diseases (IBD), inflammatory syndrome (IBS), as well as other chronic and autoimmune diseases (i.e. ...
... The third dose of HAMLET after four days (time= 3). The fourth dose after four days (time = 4) -Stools collected before the fourth dose [38,[41][42][43][44][45][46][47]. Of relevance is that the expression of the sensibility and the resistance is influenced by the selective medium culture. ...
... By another hand, the diversity and the abundance of the gut microbiota imply a role in the human health [38] in the homeostasis of nutrients metabolism, and gut immunity [39;40]. The gut Figure 1A: The gut microbiota composition under normal conditions. ...
... Firm cutes are mainly Gram-positive bacteria, and mostly Clostridium (95%) and 5% are conformed by Lactobacillus, Bacillus, Eritrococcus and Ruminococcus. While the Bacteroidetes usually are Gram negative bacteria, such as Bacteroidetes and Prevotella [38]. Any disbalance onto the gut microbiota composition leads to disease in dysbiosis, which has been associated to inflammatory diseases such as, inflammatory diseases (IBD), inflammatory syndrome (IBS), as well as other chronic and autoimmune diseases (i.e. ...
... The third dose of HAMLET after four days (time= 3). The fourth dose after four days (time = 4) -Stools collected before the fourth dose [38,[41][42][43][44][45][46][47]. Of relevance is that the expression of the sensibility and the resistance is influenced by the selective medium culture. ...
... 6 Treatment effects have further been demonstrated in patients with nonmuscle invasive bladder cancer, where intravesical alpha1-oleate instillations have been shown to trigger rapid shedding of tumor cells and tumor cell fragments into the urine, apoptosis in the tumor and a reduction in tumor size, without evidence of drug-associated toxicity. 7 Cancerrelated gene expression was reduced by about 80% in the treated tumors, including genes associated with bladder cancer. 7 The beneficial toxicity profile suggested that alpha1-oleate might be suitable for combination therapy, to enhance the effect of established, more toxic compounds at a lower dose. ...
... 7 Cancerrelated gene expression was reduced by about 80% in the treated tumors, including genes associated with bladder cancer. 7 The beneficial toxicity profile suggested that alpha1-oleate might be suitable for combination therapy, to enhance the effect of established, more toxic compounds at a lower dose. ...
... Alpha1 peptide was N-terminally labeled with the photo-stable dye JF549 and the complex with oleic acid was generated as described. 7 A-549 cells grown on μ-slide coverslips (2 x 10 5 cells) were pretreated with labeled alpha1-oleate for 20 minutes, followed by exposure to Epirubicin for 10 minutes. Cells were fixed with 2% paraformaldehyde, mounted with ProLong Glass antifade mountant (Thermofisher, Cat no P36980) and imaged using a confocal microscope (Zeiss LSM 900) with 488 and 561 nm excitation wavelengths. ...
Article
Full-text available
Bladder cancer is common and one of the most costly cancer forms, due to a lack of curative therapies. Recently, clinical safety and efficacy of the alpha1‐oleate complex was demonstrated in a placebo‐controlled study of nonmuscle invasive bladder cancer. Our study investigated if long‐term therapeutic efficacy is improved by repeated treatment cycles and by combining alpha1‐oleate with low‐dose chemotherapy. Rapidly growing bladder tumors were treated by intravesical instillation of alpha1‐oleate, Epirubicin or Mitomycin C alone or in combination. One treatment cycle arrested tumor growth, with a protective effect lasting at least 4 weeks in mice receiving 8.5 mM of alpha1‐oleate alone or 1.7 mM of alpha‐oleate combined with Epirubicin or Mitomycin C. Repeated treatment cycles extended protection, defined by a lack of bladder pathology and a virtual absence of bladder cancer‐specific gene expression. Synergy with Epirubicin was detected at the lower alpha1‐oleate concentration and in vitro, alpha1‐oleate was shown to enhance the uptake and nuclear translocation of Epirubicin, by tumor cells. Effects at the chromatin level affecting cell proliferation were further suggested by reduced BrdU incorporation. In addition, alpha1‐oleate triggered DNA fragmentation, defined by the TUNEL assay. The results suggest that bladder cancer development may be prevented long‐term in the murine model, by alpha1‐oleate alone or in combination with low‐dose Epirubicin. In addition, the combination of alpha1‐oleate and Epirubicin reduced the size of established tumors. Exploring these potent preventive and therapeutic effects will be of immediate interest in patients with bladder cancer.
... The colony number was significantly reduced in the BAMLET-treated cells (Fig. 7e,f). The concentrations of BAMLET (7, 14 and 21 µM) were selected based on extensive previous in vitro studies of HAMLET and BAMLET induced cell death with a three-hour protocol 25,28,29 . No tumoricidal effects were observed in cells exposed to 1, 3 or 5 µM of BAMLET. ...
Article
Full-text available
Though new targeted therapies for colorectal cancer, which progresses from local intestinal tumors to metastatic disease, are being developed, tumor specificity remains an important problem, and side effects a major concern. Here, we show that the protein-fatty acid complex BAMLET (bovine alpha-lactalbumin made lethal to tumor cells) can act as a peroral treatment for colorectal cancer. ApcMin/+ mice, which carry mutations relevant to hereditary and sporadic human colorectal cancer, that received BAMLET in the drinking water showed long-term protection against tumor development and decreased expression of tumor growth-, migration-, metastasis- and angiogenesis-related genes. BAMLET treatment via drinking water inhibited the Wnt/β-catenin and PD-1 signaling pathways and prolonged survival without evidence of toxicity. Systemic disease in the lungs, livers, spleens, and kidneys, which accompanied tumor progression, was inhibited by BAMLET treatment. The metabolic response to BAMLET included carbohydrate and lipid metabolism, which were inhibited in tumor prone ApcMin/+ mice and weakly regulated in C57BL/6 mice, suggesting potential health benefits of peroral BAMLET administration in addition to the potent antitumor effects. Together, these findings suggest that BAMLET administration in the drinking water maintains antitumor pressure by removing emergent cancer cells and reprogramming gene expression in intestinal and extra-intestinal tissues.
... Numerous in vitro models [3,4], animal models of cancer [5][6][7], and clinical trials [8,9] have demonstrated the antitumor activity of HAMLET. A synthetic version of HAMLET, produced by Swedish HAMLET Pharma and called Alpha1H, is currently being used in a clinical study [10]. Since the first discovery of the complex formed in breast milk by α-LA and OA, about ten proteins capable of forming HAMLETlike complexes have been discovered, all of which efficiently destruct tumor cells [11][12][13][14][15]. Lactoferrin (LF) is one such protein [16,17]. ...
Article
Full-text available
Our previous study showed that not only bovine lactoferrin (LF), the protein of milk and neutrophils, but also the human species forms complexes with oleic acid (OA) that inhibit tumor growth. Repeated injections of human LF in complex with OA (LF/8OA) to hepatoma-carrying mice decelerated tumor growth and increased animals’ longevity. However, whether the effect of the LF/8OA complex is directed exclusively against malignant cells was not studied. Hence, its effect on normal blood cells was assayed, along with its possible modulation of ceruloplasmin (CP), the preferred partner of LF among plasma proteins. The complex LF/8OA (6 μM) caused hemolysis, unlike LF alone or BSA/8OA (250 μM). The activation of neutrophils with exocytosis of myeloperoxidase (MPO), a potent oxidant, was induced by 1 μM LF/8OA, whereas BSA/8OA had a similar effect at a concentration increased by an order. The egress of heme-containing proteins, i.e., MPO and hemoglobin, from blood cells affected by LF/8OA was followed by a pronounced oxidative/halogenating stress. CP, which is the natural inhibitor of MPO, added at a concentration of 2 mol per 1 mol of LF/8OA abrogated its cytotoxic effect. It seems likely that CP can be used effectively in regulating the LF/8OA complex’s antitumor activity.
... A group of scientists from Lund University has recently published the first HAMLET data on a single-center, placebo-controlled, double-blinded randomized phase I/II interventional clinical trial of non-muscle invasive bladder cancer. Researchers concluded that intravesical inoculation of alpha1-oleate was safe and effective in patients with bladder cancer (Brisuda et al. 2021). After this successful trial, the Lund University group shared other trial ideas of having this proteolipid compound in drinking water as prevention. ...
Article
Full-text available
Purpose Treatment of advanced colorectal cancer (CRC) depends on the correct selection of personalized strategies. HAMLET (Human Alpha-lactalbumin Made LEthal to Tumor cells) is a natural proteolipid milk compound that might serve as a novel cancer prevention and therapy candidate. Our purpose was to investigate HAMLET effect on viability, death pathway and mitochondrial bioenergetics of CRC cells with different KRAS/BRAF mutational status in vitro. Methods We treated three cell lines (Caco-2, LoVo, WiDr) with HAMLET to evaluate cell metabolic activity and viability, flow cytometry of apoptotic and necrotic cells, pro- and anti-apoptotic genes, and protein expressions. Mitochondrial respiration (oxygen consumption) rate was recorded by high-resolution respirometry system Oxygraph-2 k. Results The HAMLET complex was cytotoxic to all investigated CRC cell lines and this effect is irreversible. Flow cytometry revealed that HAMLET induces necrotic cell death with a slight increase in an apoptotic cell population. WiDr cell metabolism, clonogenicity, necrosis/apoptosis level, and mitochondrial respiration were affected significantly less than other cells. Conclusion HAMLET exhibits irreversible cytotoxicity on human CRC cells in a dose-dependent manner, leading to necrotic cell death and inhibiting the extrinsic apoptosis pathway. BRAF-mutant cell line is more resistant than other type lines. HAMLET decreased mitochondrial respiration and ATP synthesis in CaCo-2 and LoVo cell lines but did not affect WiDr cells’ respiration. Pretreatment of cancer cells with HAMLET has no impact on mitochondrial outer and inner membrane permeability.
Article
Full-text available
Background The molecular content of urine is defined by filtration in the kidneys and by local release from tissues lining the urinary tract. Pathological processes and different therapies change the molecular composition of urine and a variety of markers have been analyzed in patients with bladder cancer. The response to BCG immunotherapy and chemotherapy has been extensively studied and elevated urine concentrations of IL‐1RA, IFN‐α, IFN‐γ TNF‐α, and IL‐17 have been associated with improved outcome. Methods In this study, the host response to intravesical alpha 1‐oleate treatment was characterized in patients with non‐muscle invasive bladder cancer by proteomic and transcriptomic analysis. Results Proteomic profiling detected a significant increase in multiple cytokines in the treatment group compared to placebo. The innate immune response was strongly activated, including IL‐1RA and pro‐inflammatory cytokines in the IL‐1 family (IL‐1α, IL‐1β, IL‐33), chemokines (MIP‐1α, IL‐8), and interferons (IFN‐α2, IFN‐γ). Adaptive immune mediators included IL‐12, Granzyme B, CD40, PD‐L1, and IL‐17D, suggesting broad effects of alpha 1‐oleate treatment on the tumor tissues. Conclusions The cytokine response profile in alpha 1‐oleate treated patients was similar to that reported in BCG treated patients, suggesting a significant overlap. A reduction in protein levels at the end of treatment coincided with inhibition of cancer‐related gene expression in tissue biopsies, consistent with a positive treatment effect. Thus, in addition to killing tumor cells and inducing cell detachment, alpha 1‐oleate is shown to activate a broad immune response with a protective potential.
Article
Full-text available
As the most aggressive subtype, triple‐negative breast cancer (TNBC) without definitive targets represents a high probability of metastasis. Nevertheless, the presence of high‐level tumor‐infiltrating lymphocytes in the TNBC tumor microenvironment (TME) suggests patients may benefit from immunotherapy. In this study, bovine α‐lactalbumin coupled with oleic acid that forms tumoricidal lipid–protein complex (BAMLET) is electrostatically stabilized on the surface of amino‐functionalized mesoporous silica nanoparticles (MSN‐NH2). It is found that MSN‐NH2 may cause partial conformational changes of immobilized proteins due to electrostatic interactions. Therefore, BAMLET covered MSN‐NH2 (BMSN) as a drug cargo that can not only induce selectively oncolytic effect but kill tumor cells with rapid kinetics. Moreover, the hemolytic activity of BMSN conjugated with pre‐formed serum protein corona is largely reduced which may avoid the poor hemocompatibility caused by BAMLET. Furthermore, the in vivo study indicates that the combinational use of drug‐loaded BMSN and immune checkpoint small‐molecule inhibitor can efficiently inhibit the growth of solid TNBC tumors and activate immune response by sufficient infiltration of immune cells in the TME and prevent the seeding of circulating tumor cells. Therefore, the constructed BMSN provides an integrated nanotherapeutics and a nanoadjuvant with superior anticancer performance to combat TNBC.
Article
Bladder cancer (BC) is featured as the second most common malignancy of the urinary tract worldwide with few treatments leading to high incidence and mortality. It stayed a virtually intractable disease, and efforts to identify innovative and effective therapies are urgently needed. At present, more and more evidence shows the importance of non-coding RNA (ncRNA) for disease-related study, diagnosis, and treatment of diverse types of malignancies. Recent evidence suggests that dysregulated functions of ncRNAs are closely associated with the pathogenesis of numerous cancers including BC. The detailed mechanisms underlying the dysregulated role of ncRNAs in cancer progression are still not fully understood. This review mainly summarizes recent findings on regulatory mechanisms of the ncRNAs, long non-coding RNAs, microRNAs, and circular RNAs, in cancer progression or suppression and focuses on the predictive values of ncRNAs-related signatures in BC clinical outcomes. A deeper understanding of the ncRNA interactive network could be compelling framework for developing biomarker-guided clinical trials.
Article
Full-text available
Background: Human milk is recognized as an ideal food for newborns and infants owing to the presence of various nutritive factors, including healthy bacteria. Aim/Objective: This review aimed to understand the effects of human milk microbiota in both the prevention of disease and the health of infants. Methods: Data were obtained from PubMed, Scopus, Web of Science, clinical trial registries, Dergipark, and Türk Atıf Dizini up to February 2023 without language restrictions. Results: It is considered that the first human milk microbiota ingested by the newborn creates the initial microbiome of the gut system, which in turn influences the development and maturation of immunity. Bacteria present in human milk modulate the anti-inflammatory response by releasing certain cytokines, protecting the newborn against certain infections. Therefore, certain bacterial strains isolated from human milk could serve as potential probiotics for various therapeutic applications. Conclusions: In this review, the origin and significance of human milk bacteria have been highlighted along with certain factors influencing the composition of human milk microbiota. In addition, it also summarizes the health benefits of human milk as a protective agent against certain diseases and ailments.
Article
Full-text available
Potent chemotherapeutic agents are required to counteract the aggressive behavior of cancer cells and patients often experience severe side effects, due to tissue toxicity. Our study addresses if a better balance between efficacy and toxicity can be attained using the tumoricidal complex alpha1‐oleate, formed by a synthetic, alpha‐helical peptide comprising the N‐terminal 39 amino acids of alpha‐lactalbumin and the fatty acid oleic acid. Bladder cancer was established, by intravesical instillation of MB49 cells on day 0 and the treatment group received five instillations of alpha1‐oleate (1.7‐17 mM) on days 3 to 11. A dose‐dependent reduction in tumor size, bladder size and bladder weight was recorded in the alpha1‐oleate treated group, compared to sham‐treated mice. Tumor markers Ki‐67, Cyclin D1 and VEGF were inhibited in a dose‐dependent manner, as was the expression of cancer‐related genes. Remarkably, toxicity for healthy tissue was not detected in alpha1‐oleate‐treated, tumor‐bearing mice or healthy mice or rabbits, challenged with increasing doses of the active complex. The results define a dose‐dependent therapeutic effect of alpha1‐oleate in a murine bladder cancer model.
Article
Full-text available
Context Currently, there is no standard of care for patients with non–muscle-invasive bladder cancer (NMIBC) who recur despite bacillus Calmette-Guerin (BCG) therapy. Although radical cystectomy is recommended, many patients decline to undergo or are ineligible to receive it. Multiple agents are being investigated for use in this patient population. Objective To systematically synthesize and describe the efficacy and safety of current and emerging treatments for NMIBC patients after treatment with BCG. Evidence acquisition A systematic literature search of MEDLINE, Embase, and the Cochrane Controlled Register of Trials (period limited to January 2007–June 2019) was performed. Abstracts and presentations from major conference proceedings were also reviewed. Randomized controlled trials were assessed using the Cochrane risk of bias tool. Data for single-arm trials were pooled using a random-effect meta-analysis with the proportions approach. Trials were grouped based on the minimum number of prior BCG courses required before enrollment and further stratified based on the proportion of patients with carcinoma in situ (CIS). Evidence synthesis Thirty publications were identified with data from 23 trials for meta-analysis, of which 17 were single arm. Efficacy and safety outcomes varied widely across studies. Heterogeneity across trials was reduced in subgroup analyses. The pooled 12-mo response rates were 24% (95% confidence interval [CI]: 16–32%) for trials with two or more prior BCG courses and 36% (95% CI: 25–47%) for those with one or more prior BCG courses. In a subgroup analysis, inclusion of ≥50% of patients with CIS was associated with a lower response. Conclusions The variability in efficacy and safety outcomes highlights the need for consistent endpoint reporting and patient population definitions. With promising emerging treatments currently in development, efficacious and safe therapeutic options are urgently needed for this difficult-to-treat patient population. Patient summary We examined the efficacy and safety outcomes of treatments for non–muscle-invasive bladder cancer after bacillus Calmette-Guerin therapy. Outcomes varied across studies and patient populations, but emerging treatments currently in development show promising efficacy.
Article
Full-text available
Background: Nonmuscle-invasive bladder cancer (NMIBC) is the most common form of bladder cancer, with high rates of disease recurrence and progression. Current treatment for high-risk NMIBC involves Bacillus Calmette-Guérin (BCG) therapy, but treatment options are limited for patients with recurrent or BCG-unresponsive disease. Aberrant programmed death 1 signaling has been implicated in BCG resistance and bladder cancer recurrence and progression, and pembrolizumab has shown efficacy in patients with BCG-unresponsive high-risk NMIBC. Aim: To describe the rationale and design for the randomized, comparator-controlled Phase III KEYNOTE-676 study, which will evaluate the efficacy and safety of pembrolizumab in combination with BCG in patients with persistent/recurrent high-risk NMIBC after BCG induction therapy. Trial registration number: NCT03711032
Article
Full-text available
Background Adjuvant intravesical instillations with bacillus Calmette-Guérin (BCG) is the recommended treatment option for patients with intermediate-and high-risk non-muscle invasive bladder cancer (NMIBC). Despite adequate BCG treatment, a large proportion of patients experience a recurrence. Although radical cystectomy is the gold standard for BCG unresponsive NMIBC, some patients are unfit or unwilling to consider this option. Objective To assess the effectiveness of Hyperthermic IntraVEsical Chemotherapy (HIVEC®) in BCG unresponsive NMIBC patients. Methods A post-hoc analysis was conducted of prospectively included intermediate-and high-risk NMIBC patients who were planned to receive HIVEC® treatment between October 2014 and November 2017. For the present analysis, only patients who met the BCG unresponsive definition were included. Patients were followed by cystoscopy and cytology every 3 months and a CT-urography scan yearly. The primary outcome was the disease-free survival (DFS). The Common Terminology Criteria for Adverse Events (CTCAE) was used to assess side-effects. Results The study population consisted of 55 BCG unresponsive NMIBC patients of whom 52 underwent≥5 HIVEC® treatments. The median age and follow-up were 73 years and 14.0 months (IQR 7.6 – 24.6). The median DFS was 17.7 months (SE 6.72) and progression occurred in four patients. The 1-year cumulative incidence rate of disease recurrence/progression was 53%. Two patients experienced severe side-effects (CTCAE≥3). Conclusions HIVEC® seems a valid treatment option for BCG unresponsive NMIBC patients. We report a median DFS of 17.7 months (SE 6.72), potentially avoiding or postponing the need for radical surgery in a proportion of these patients.
Article
Full-text available
Peptides and proteins have been found to possess an inherent tendency to convert from their native functional states into intractable amyloid aggregates. This phenomenon is associated with a range of increasingly common human disorders, including Alzheimer and Parkinson diseases, type II diabetes, and a number of systemic amyloidoses. In this review, we describe this field of science with particular reference to the advances that have been made over the last decade in our understanding of its fundamental nature and consequences. We list the proteins that are known to be deposited as amyloid or other type of aggregates in human tissues and the disorders with which they are associated, as well as the proteins that exploit the amyloid motif to play specific functional roles in humans. In addition, we summarize the genetic factors that have provided insight into the mechanisms of disease onset. We describe recent advances in our knowledge of the structures of amyloid fibrils and their oligomeric precursors and of the mechanisms by which they are formed and proliferate to generate cellular dysfunction. We show evidence that a complex proteostasis network actively combats protein aggregation and that such an efficient system can fail in some circumstances and give rise to disease. Finally, we anticipate the development of novel therapeutic strategies with which to prevent or treat these highly debilitating and currently incurable conditions. Expected final online publication date for the Annual Review of Biochemistry Volume 86 is June 20, 2017. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
Background BCG is the most effective therapy for high-risk non-muscle-invasive bladder cancer. Nadofaragene firadenovec (also known as rAd-IFNa/Syn3) is a replication-deficient recombinant adenovirus that delivers human interferon alfa-2b cDNA into the bladder epithelium, and a novel intravesical therapy for BCG-unresponsive non-muscle-invasive bladder cancer. We aimed to evaluate its efficacy in patients with BCG-unresponsive non-muscle-invasive bladder cancer. Methods In this phase 3, multicentre, open-label, repeat-dose study done in 33 centres (hospitals and clinics) in the USA, we recruited patients aged 18 years or older, with BCG-unresponsive non-muscle-invasive bladder cancer and an Eastern Cooperative Oncology Group status of 2 or less. Patients were excluded if they had upper urinary tract disease, urothelial carcinoma within the prostatic urethra, lymphovascular invasion, micropapillary disease, or hydronephrosis. Eligible patients received a single intravesical 75 mL dose of nadofaragene firadenovec (3 × 10¹¹ viral particles per mL). Repeat dosing at months 3, 6, and 9 was done in the absence of high-grade recurrence. The primary endpoint was complete response at any time in patients with carcinoma in situ (with or without a high-grade Ta or T1 tumour). The null hypothesis specified a complete response rate of less than 27% in this cohort. Efficacy analyses were done on the per-protocol population, to include only patients strictly meeting the BCG-unresponsive definition. Safety analyses were done in all patients who received at least one dose of treatment. The study is ongoing, with a planned 4-year treatment and monitoring phase. This study is registered with ClinicalTrials.gov, NCT02773849. Findings Between Sept 19, 2016, and May 24, 2019, 198 patients were assessed for eligibility. 41 patients were excluded, and 157 were enrolled and received at least one dose of the study drug. Six patients did not meet the definition of BCG-unresponsive non-muscle-invasive bladder cancer and were therefore excluded from efficacy analyses; the remaining 151 patients were included in the per-protocol efficacy analyses. 55 (53·4%) of 103 patients with carcinoma in situ (with or without a high-grade Ta or T1 tumour) had a complete response within 3 months of the first dose and this response was maintained in 25 (45·5%) of 55 patients at 12 months. Micturition urgency was the most common grade 3–4 study drug-related adverse event (two [1%] of 157 patients, both grade 3), and there were no treatment-related deaths. Interpretation Intravesical nadofaragene firadenovec was efficacious, with a favourable benefit:risk ratio, in patients with BCG-unresponsive non-muscle-invasive bladder cancer. This represents a novel treatment option in a therapeutically challenging disease state. Funding FKD Therapies Oy.
Article
Background Between 2013 and 2016, global production of bacillus Calmette-Guérin (BCG) was dramatically reduced due to the collapse of the factory producing BCG Connaught. Objective To evaluate the clinical and economic impact of BCG shortage on a cohort of non–muscle-invasive bladder cancer (NMIBC) patients treated during the period of restricted supply. Design, setting and participants This retrospective, before and after, cost-consequence study included patients with intermediate- and high-risk NMIBC. Those resected between November 2011 and September 2013 (control group) were compared with those resected between October 2013 and December 2016 (study group). Outcome measurements and statistical analysis The primary endpoint was the rate of tumor recurrence from 30 d after transurethral resection to the end of follow-up at 24 mo; the secondary endpoints included the average cost of primary treatment, average cost of treatment of recurrence, and excess cost due to BCG shortage per patient. Results and limitations A total of 402 patients were included: 191 in the control group and 211 in the study group. The rate of recurrence at 24 mo was significantly higher in the study group than in the control group (46.9% vs 16.2%; relative risk: 0.7, 95% confidence interval [0.60; 0.82]; p < 0.001). The increased cost due to the decrease in BCG production was estimated to be €783 per patient with a new diagnosis of NMIBC during the period of restricted supply. This is a retrospective analysis at the level of our unit. A more precise evaluation would require a study of a larger cohort of patients. Conclusions The shortage of BCG between October 2013 and December 2016 had a significant medical and economic impact; there was an increased rate of bladder cancer recurrence, and the total cost of care for intermediate- and high-risk NMIBC was higher. Patient summary In this report, we analyzed the medical and economic impact of bacillus Calmette-Guérin (BCG) shortage that occurred between 2013 and 2016. We found a significant increase of bladder cancer recurrence and progression, and an increase in the number of patients who had to be treated by cystectomy. BCG shortage also had a significant impact on the total cost. Since there are no alternatives to BCG for high-risk non–muscle-invasive bladder cancer patients, BCG production has to be maintained by any means.
Book
This book describes a novel and proven approach to cytologically classify urinary samples for the detection of bladder cancer and lesions of the upper urinary tract. The new method is based on the collective experience of knowledgeable cytopathologists who have tested the terminology within their own laboratories for reproducibility and predictability of neoplasms of the urinary tract. Accompanying the written criteria for each diagnostic category are meticulously photographed exemplars of the cellular features, with cogently annotated descriptions of the photographs. The book thereby performs as an atlas for microscopists involved in diagnostic cytopathology at all levels of their education. Included in the targeted readership are experienced pathologists, cytotechnologists, and students of both professional groups. The new terminology also considers the clinical aspects of patient management. Written by experts in the field who convened at the 18th International Congress of Cytology in Paris, The Paris System for Reporting Urinary Cytology presents a global standard for reporting and a new philosophic approach that maximizes the strengths of detecting the potentially lethal high grade lesions by urinary cytology, and recognizes without apology the inability to reliably detect the low grade lesions in urinary cytology. The Concept has been endorsed by the American Society Of Cytopathology, and the International Academy of Cytology.
Article
Background: Survival in patients with bladder cancer has only moderately improved over the past 2 decades. A potential reason for this is nonadherence to clinical guidelines and best practice, leading to wide variations in care. Common quality indicators (QIs) are needed to quantify adherence to best practice and provide data for benchmarking and quality improvement. Objective: To produce an evidence- and consensus-based list of QIs for the management of bladder cancer. Methods: A modified Delphi method was used to develop the indicator list. Candidate indicators were extracted from the literature and rated by a 27-member Canadian expert panel in several rounds until consensus was reached on the final list of indicators. In rounds with numeric ratings, a frequency analysis was performed. Results: A total of 86 indicators were rated, 52 extracted from the literature and 34 suggested by the panel. After iterative rounds of ratings and discussion, a final list of 60 QIs spanning several disciplines and phases of the cancer care continuum was developed. Conclusions: This is the first study to comprehensively produce common QIs representing structure, process, and outcome measures in bladder cancer management. Though developed in Canada, these indicators can be used in other countries with slight modifications to track performance and improve care.
Article
Context: Bladder cancer has become a common cancer globally, with an estimated 430 000 new cases diagnosed in 2012. Objective: We examine the most recent global bladder cancer incidence and mortality patterns and trends, the current understanding of the aetiology of the disease, and specific issues that may influence the registration and reporting of bladder cancer. Evidence acquisition: Global bladder cancer incidence and mortality statistics are based on data from the International Agency for Research on Cancer and the World Health Organisation (Cancer Incidence in Five Continents, GLOBOCAN, and the World Health Organisation Mortality). Evidence synthesis: Bladder cancer ranks as the ninth most frequently-diagnosed cancer worldwide, with the highest incidence rates observed in men in Southern and Western Europe, North America, as well in certain countries in Northern Africa or Western Asia. Incidence rates are consistently lower in women than men, although sex differences varied greatly between countries. Diverging incidence trends were also observed by sex in many countries, with stabilising or declining rates in men but some increasing trends seen for women. Bladder cancer ranks 13th in terms of deaths ranks, with mortality rates decreasing particularly in the most developed countries; the exceptions are countries undergoing rapid economic transition, including in Central and South America, some central, southern, and eastern European countries, and the Baltic countries. Conclusions: The observed patterns and trends of bladder cancer incidence worldwide appear to reflect the prevalence of tobacco smoking, although infection with Schistosoma haematobium and other risk factors are major causes in selected populations. Differences in coding and registration practices need to be considered when comparing bladder cancer statistics geographically or over time. Patient summary: The main risk factor for bladder cancer is tobacco smoking. The observed patterns and trends of bladder cancer incidence worldwide appear to reflect the prevalence of tobacco smoking.