Skip to main content
Erschienen in: BMC Hematology 1/2017

Open Access 01.12.2017 | Research article

A novel fibrinogen mutation: FGA g. 3057 C > T (p. Arg104 > Cys) impairs fibrinogen secretion

verfasst von: R. Marchi, M. Linares, H. Rojas, A. Ruiz-Sáez, M. Meyer, A. Casini, S.O. Brennan

Erschienen in: BMC Hematology | Ausgabe 1/2017

Abstract

Background

Abnormal fibrinogens can be caused by clinically silent hereditary mutations. A new case was detected accidentally in an 11-year-old girl when routine pre-operative coagulation tests were performed for nasal turbinate surgery.

Methods

The fibrinogen genes FGA, FGG and FGB were sequenced using standard protocols. The kinetics of fibrin formation were followed by turbidity at 350 nm. Purified fibrinogen was incubated with plasmin, and the degradation products analyzed by SDS/PAGE. The formation of fibrinogen-albumin complexes was analyzed by immunobloting. Fibrin structure was examined in a Nikon Eclipse TE 2000-U laser microscope. Secretion of the variant protein was analyzed directly by reverse phase-electrospray time of flight-mass spectrometry (TOF-MS).

Results

DNA sequencing revealed a novel heterozygous g. 3057 C > T mutation in the FGA that predicts a p. Arg104 > Cys substitution, in the proband and her father. Both patients were asymptomatic with low functional and antigen fibrinogen concentrations. The proband’s plasma fibrinogen polymerization was almost normal, with a 12% decrease in the final turbidity, while, the father’s fibrin formation had a diminished slope and final turbidity (2.5× and 40%, respectively). Aα Arg104 is located at a plasmin cleavage site in the coiled-coil region of fibrinogen. However, the father’s fibrinogen plasmin degradation was normal. Although the exchanged Cys introduces an unpaired –SH, immunoblotting showed no fibrinogen-albumin complexes. Furthermore, the plasma clot structure observed by confocal microscopy appeared almost normal. TOF-MS showed that the variant Aα chain was underrepresented in plasma and made up only about 25% of the total.

Conclusions

The low expression of the Aα Arg104 > Cys chain in circulation could account for the observed hypodysfibrinogenemia.
Hinweise

Electronic supplementary material

The online version of this article (https://​doi.​org/​10.​1186/​s12878-017-0086-8) contains supplementary material, which is available to authorized users.
Abkürzungen
a.u.
Arbitrary unit
aPTT
Activated partial thromboplastin time
DAB
3, 3’diaminobenzidine
DTT
Dithiothreitrol
EDTA
Ethylenediaminetetraacetatic acid
EM
Elastic modulus
Fg
Fibrinogen
FXIIIa
Activated factor XIII
HAS
Hemostasis analyzer system
Ks
Permeation constant
MaxAbs
Maximum absorbance
OD
Optical density
PCR
Polymerase chain reaction
PPP
Platelet poor plasma
PT
Prothrombin time
SDS/PAGE
Sodium dodecyl sulfate polyacrylamide gel electrophoresis
TBS
Tris buffered saline
TF
Tissue factor
TOF-MS
Reverse phase-electrospray time of flight-mass spectrometry
tPA
Tissue type plasminogen activator
TT
Thrombin time

Background

Fibrinogen is the central protein of blood coagulation. Once the coagulation cascade is initiated, thrombin is formed and catalyzes the conversion of fibrinogen into soluble fibrin monomers that polymerize spontaneously, forming a three-dimensional network that becomes further stabilized by activated factor XIII (FXIIIa). Polymerization is initiated by cleavage of the A and B peptides from the N-terminal of the Aα and Bβ chains [1]. Fibrinogen is a 340 kDa glycoprotein synthetized in the liver and normally circulates in plasma at 160–400 mg/dl [2]. It is composed of two sets of three different polypeptides chains (Aα, Bβ,γ)2, arranged in three nodules: the N-terminal of the six chains converge at the center forming the globular E region. A coiled coil of all the three chains extends from each side of the E domain to connect with the outer D domains, which form from the C-terminal region of the Bβ and γ chains. The coiled coil is delineated by two disulfide rings and its central region has a kink in its structure that acts as the primary attack site for plasmin [3].
Inherited fibrinogen disorders can be quantitative (Type I; absence or decreased level of circulating fibrinogen, afibrinogenemia and hypofibrinogenemia, respectively) or qualitative (Type 2; normal or decreased antigenic levels and low fibrinogen activity, dysfibrinogenemia and hypodysfibrinogenemia, respectively) [2, 4].
Dysfibrinogenemia is caused by structural abnormalities that can be inherited (congenital) or acquired [5]. Inherited dysfibrinogenemia is caused by mutations in the coding region of the fibrinogen Aα, Bβ or γ genes and the majority of cases result from heterozygous missense mutations [4]. The prevalence of inherited dysfibrinogenemia in the general population is unknown [5]. The pattern of inheritance is autosomal dominant, and 55% of the patients are asymptomatic, while 25% develop bleeding or thrombosis. Hypodysfibrinogenemia has features of both hypo- and dysfibrinogenemia: the reduced circulating fibrinogen levels confers a hypofibrinogenemia phenotype, and the expression of the mutation that alters functionality, a dysfibrinogenemic phenotype [6]. As with dysfibrinogenemia, in hypodysfibrinogenaemia bleeding extends from mild to moderate, but individuals are more predisposed to thrombosis.
The diagnosis of qualitative fibrinogen disorders is done by the measurement of standard clotting times, whose sensitivity depends upon the methods, reagents and coagulometers used [4]. Usually the thrombin time is prolonged, although with some variants it may be normal [5].
During fibrinogen synthesis, each newly synthetized chain is independently translocated into the endoplasmic reticulum (ER) where chaperones assist in the assembly and folding processes. Molecules are assembled in a step wise manner in the lumen of the ER: first two-chain Aα-γ and Bβ-γ complex are formed. These complexes recruit a Bβ or Aα chain respectively and form half-molecules (Aα-Bβ-γ)1, which in the last step dimerize through N-terminal disulphide bridges to form (Aα-Bβ-γ)2 hexamers [7]. Several studies performed in recombinant systems, using deletion and substitution mutants, indicate that an intact coiled-coil and correct inter- and intrachain disulphide bonds are needed for final molecular assembly [810]. In afibrinogenemia mutated molecules are usually absent from circulation. However, in hypofibrinogenemia or hypodysfibrinogenemia if the mutation impairs assembly variant molecules may be secreted, but underrepresented in plasma.
Here we describe a new variant with an Aα Arg104 → Cys mutation in the coiled-coil region that we have named fibrinogen Caracas IX.

Methods

Materials

Bovine thrombin was from Enzyme Research Laboratories (South Bend, IN). Lysine-sepharose 4B was from GE Healthcare (Piscataway, NJ). Pefabloc® SC (4- (2-Aminoethyl) benzenesulfonyl fluoride hydrochloride) was from Fluka, Sigma-Aldrich (Buchs, Switzerland). Tissue-type plasminogen activator and plasmin were from American Diagnostica (Stamford, CT). Human albumin and benzamidine were from Sigma Chemical Company (St Louis, MO). Albumin antibody conjugated to peroxidase was from Dako Corporation (Carpinteria, CA). The substrate 3,3′-diaminobenzidine (DAB) was from Thermo Scientific (Rockford, IL). The LabTek chambers and Alexa Fluor 488 were purchased from Invitrogen, Nalge Nunc International (Rochester, NY).

Blood collection

Blood was collected in citrate (1 volume of 0.13 M trisodium citrate and 9 volumes of blood), the first 3 ml of blood discarded, and centrifuged twice at 2000 g for 10 min. The platelet poor plasma obtained (PPP) was supplemented with benzamidine 10 mM (final concentration), except the plasma to be used for fibrinolysis experiments, aliquoted and kept at −80 °C until use. Routine coagulation tests were performed with citrated plasma on coagulation analyzer STA Compact®, Stago, France. Fibrinogen level was determined by Clauss (Laboratoire Stago, Asnière, France) and clot weight method [11]. Antigenic fibrinogen concentration was measured by a latex immunoassay (Liaphen Fibrinogen, Hyphen BioMed, France).

Mutation analysis

Blood was collected in 0.5 M ethylenediaminetetraacetate tetrasodium salt (EDTA Na4) (50:1). Genomic DNA was isolated using the Invisorb Spin Blood Mini Kit (Invitek GmbH, Berlin, Germany) according to the manufacturer’s protocol. Sequences comprising all exons and exon-intron boundaries from the three fibrinogen genes: FGA, FGB, and FGG were amplified by the polymerase chain reaction (PCR) according to standard protocols. After purification of the PCR products using the Invisorb Spin PCRapid Kit® (Invitek, Berlin, Germany), direct DNA cycle sequencing was performed, applying the Big Dye kit from Applied Biosystems (Foster City, CA), according to the manufacturer’s recommendations.

Fibrinogen purification

Fibrinogen was purified essentially as described elsewhere with modifications [12]. Plasma samples were thawed and supplemented with 1 mM Pefabloc® and 5 mM EDTA (final concentrations). Plasma was depleted of plasminogen by passing through a lysine-sepharose 4B column, and then fibrinogen was purified by precipitation (×3) with 25%-saturation ammonium sulphate, pH 7.5. This fibrinogen fraction also contained co-purifying fibronectin, factor XIII and vW factor. The precipitate was dissolved in 0.3 M NaCl, dialyzed against the same solution, and stored at −80 °C until used.
The integrity of the purified fibrinogen was analyzed by sodium dodecylsulfate-polyacilamide gel electrophoresis (SDS-PAGE) on 8% gel. The coagulability of the purified fibrinogen was 96% and 93% and the yield 43 and 27%, control and patient, respectively.

Fibrinogen degradation

Fibrinogen was incubated with plasmin as described [13] with minor modifications. Purified fibrinogen (0.9 mg/ml, in TBS) was incubated with plasmin (18 μg/ml, in TBS) in the presence of 1 mM CaCl2 or 5 mM EDTA at 37 °C at different incubation times (15, 30 min and 4 h), quenched with 2% SDS-DTT (v:v) sample buffer and immediately boiled. The zero time sample contained no plasmin. The fibrinogen degradation products were analyzed in a 6% gel SDS-PAGE under non reducing conditions.

Western blotting

In order to detect fibrinogen-albumin complexes, Western blotting was performed under non reduced conditions essentially as described [14]. Briefly, purified fibrinogens (5 μg) and human albumin (5 μg) were loaded in a 5% gel SDS-PAGE, and electroblotted onto nitrocellulose [15]. The membrane was incubated for 2 h with anti-human albumin antibody conjugated to peroxidase (1:1000). The cross-reacting bands were detected with 0.6% 3, 3’diaminobenzidine (DAB), 3% cobalt chloride and 3% hydrogen peroxide.

Activated factor XIII (FXIIIa) fibrin cross-linking

The kinetics of fibrin cross-linking was examined essentially as described elsewhere [14]. Fibrin was cross-linked by the endogenous factor XIIIa that precipitated together with fibrinogen during the purification process. Purified fibrinogen (1 mg/ml) was clotted with 1 U/ml of thrombin and 5 mM CaCl2. The reactions were quenched at different time points (0, 2, 5, 15 min and 1, 4, 24 h) with 2% SDS-DTT and analyzed in a 8% gel SDS-PAGE.

Fibrin polymerization

The kinetics of fibrin formation were studied in plasma and purified fibrinogen [16]. Briefly, 100 μl fresh plasma or 0.5 mg/ml purified fibrinogen in 50 mM Tris, 0.15 M NaCl, pH 7.4 (TBS) were dispensed in a 96-well plate. Then 10 μl of 1 unit/ml bovine thrombin - 20 mM CaCl2 (final) was added to plasma or 5 units/ml bovine thrombin and 5 mM CaCl2 to fibrinogen solution. The changes in optical density (OD) were recorded every 15 s over 1 h at 350 nm in a Tecan Infinite® M 200. The polymerizations were done in three different experiments in triplicate. The lag time (s), slope (mOD/s) and final turbidity (mOD) were calculated from each curve and averaged.

Fibrinolysis

The method was performed as described by Carter et at 2007 [17] with minor modifications. The PPP aliquots without benzamidine were used. Twenty five μl of PPP were aliquoted in a 96-well plate, then 75 μl of tPA 166 ng/ml diluted in 20 mM Hepes, 0.15 M NaCl, pH 7.4 was added. Clotting was initiated by adding 50 μl of thrombin-CaCl2 (0.03 U/ml and 9 mM, respectively).The OD changes were recorded at 350 nm every 15 s over 1.5 h in a TECAN® infinite 2 M microplate reader. The fibrinolysis was done at least three times in triplicate. The time to degrade 50% of the clot (T50%) was calculated from the time elapsed from half the value of the maximum absorbance of the polymerization to half-the value decreased of the maximum absorbance of the lysis curve branch. The rate of clot degradation (slope) was calculated in the descending part of the curve and the absolute value reported.

Direct plasma mass spectrometry

Plasma was precipitated with saturated (NH4)2SO4 (25%, final), and the precipitate washed (2×) with 25% saturated (NH4)2SO4. The pellet was dissolved in 8 M urea, 30 mM dithiothreitol, 50 mM Tris–HCl, pH 8.0 and left 3 h at 37 °C. The reduced sample was injected into an Agilent 6230 Accurate-Mass electrospray time-of-flight (TOF) mass spectrometry system [18]. A Poroshell 300SB C3 (2.1 × 75 mm) column was used with an acetonitrile gradient and profile data was collected. Multi charged spectral envelopes were deconvoluted using maximum entropy processing and BioConfirm software with an isotope width of 15 Da.

Clots biophysical characterization

In order to characterize some biophysical parameters of the clot structure, the elastic modulus, the Darcy constant (Ks), and fibrin network imaging by confocal microscopy were performed. For these experiments a healthy man with plasma fibrinogen concentration similar to the patient was chosen as a control.

Elastic modulus

The fibrin elastic modulus (EM) was measured in the hemostasis analyzer system (HAS) Hemodyne® (Richmond, VA). Briefly, 700 μL of plasma was placed in the plastic cone and incubated for 1 min at 37 °C. Then 50 μL of a thrombin - CaCl2 solution (1.3 U/ml and 25 mM final concentrations; respectively) was gently and carefully mixed in. The increase in the EM was recorded every 1 min over a 30 min period. Each sample was run in triplicate in three independent experiments. The EM (kdyne/cm2) reported corresponds to the averaged EM value that was reached at 30 min.

Permeation

Permeation through plasma clots was recorded essentially as described elsewhere [19]. The clotting conditions used were 1 U/ml of thrombin and 20 mM CaCl2 (final concentrations). The clots were left for 2 h in a moist environment at 37 °C in order to fully polymerize. The buffer percolated through the columns was TBS. Nine clots of each sample were run per experiment (n = 3) and one measurement per clot was taken.
The Darcy constant (Ks) was calculated using the following equation [20]:
$$ \mathrm{Ks}=\mathrm{QL}\upeta /\mathrm{tAP} $$
Where Q = volume of the buffer (in cm3), having a viscosity η of 0.01(poise), flowing through the column of height L (cm) and area A (cm2) in a given time (s) under a hydrostatic pressure P (dyne/cm2).

Confocal microscopy

The experiments were done essentially as described elsewhere [16]. Briefly, clots were formed inside the eight wells LabTek chambers (Invitrogen, Nalge Nunc International, Rochester, NY). Plasma samples were mixed with Alexa Fluor 488-labeled fibrinogen (10 μg/315 μl final sample volume), then clotted with thrombin-CaCl2 solution (0.3 U/ml and 20 mM, respectively, final concentration). The clots were left for 2 h in a moist environment at 37 °C in order to fully polymerize.
The fibrin clots were observed in a Nikon Eclipse TE 2000 U laser scanning confocal microscopy (LSCM), with an argon ion laser (473 nm excitation and 520/540 nm for emission). The objective used was Plan APO VC 60X water immersion with a work distance of 0.27. The acquisition pinhole was set to 60 μm. Image analyses were done as described [21]. A z-stack of 60 slice was use for construct a 3D projection of 30 μm thick (0.5 μm/slice) were done. Five image by clot (212 × 212 μm) for each experiment (control and patient) were accomplished. Two diagonal lines, a horizontal and a vertical were drawn on the volumetric image of the stack using the Olympus FV10-ASW 2.1 software for obtain the pseudocolor perfil by line. Line graphs were used to calculate density (picks/μ) and diameter of fibers (μm) with Origin Pro 8 software.

Dynamic fibrin clot growth

The spatio temporal dynamics of fibrin clot formation in real time was assessed in plasma by measuring light scattering over 30 min every 15 s using a Thrombodynamics Analyser System (HemaCore, Moscow, Russia) as previously described [21]. Briefly, plasma coagulation is activated when it is brought in contact with tissue factor coated on a plastic cuvette. The clot formation begins on the activator and propagates into the bulk of plasma in which no TF is present. Images are analyzed computationally to measure lag time, initial and stationary growth rate, size at 30 min and clot density. Based on the plots of clot size versus time, the initial velocity of clot growth is measured as the mean slope over the 2–6 min period (characterizing the VIIa-TF pathway) and the stationary velocity of clot growth is measured as the mean slope over the 15–25 min period [22].

Results

Case report

A new abnormal fibrinogen was discovered accidentally in an asymptomatic 11-year-old girl when routine pre-operative coagulation tests were performed for nasal turbinate surgery, which proceeded successfully. Her parents reported that when she was 3 years old she had surgery for stenosing tenosynovitis without any complications. Haemostasis work-up showed that the thrombin time was marginally prolonged, and that functional and antigenic fibrinogen levels were decreased without discrepancy in the proband and her father (Table 1). Neither parent report any hemostatic problems. Full DNA sequencing of FGA, FGB and FGG revealed that proband was heterozygous for a novel point mutation in FGA g. 3057 C > T that gives rise to an Aα Arg104 > Cys substitution (numbered without the signal peptide). In addition, she was found to be heterozygous for the Aα Ala312/Thr polymorphism. Targeted sequencing of the father showed he was also heterozygous for the novel Aα Arg104 > Cys mutation but his Aα Ala312/Thr status was not explored.
Table 1
Summary of coagulation tests
 
Proband
Father
Reference values
TT (s)
18.6
30.8
16–18
PT (s)
15.8
54.6
13–15
aPTT (s)
34.6
42.2
27–34
Fibrinogen (Clauss)
163
122
 
(mg/dl)
151
 
200–400
158
  
Fibrinogen antigen (mg/dl)
177
126
190–400
Factor V (%)
79
60–140
Factor II (%)
105
88
60–140
Factor VII (%)
79
69
65–145
Factor X (%)
82
67
65–130
TT thrombin time, PT prothrombin time, aPTT activated partial thromboplastin time
-: not done

Plasma mass spectrometry

Electrospray TOF MS of purified fibrinogen from the proband and her father showed normal masses and isoforms for the Bβ and γ chain components, with no evidence of any mutations (not shown). Examination of extracted Aα chain spectra from a control, who was homozygous for the AαThr312 polymorphism, showed the expected major peak at 66,136 Da corresponding to the non-phosphorylated form of the Aα chain (theoretical mass 66,132 Da) (Fig. 1). With a peak at 66,134 Da the father appeared to have one normal copy of the AαThr312 allele and one variant copy giving rise to a protein of mass 66,080 Da. This mass decrease of 54 Da was entirely consistent with a point mutation of Arg → Cys (−53 Da). Spectra shows the proband inherited this same variant (theoretical mass 66,079 Da) from her father, together with a copy of the less common AαAla312 allele from her mother. Interestingly the variant chain with the Arg → Cys substitution was underrepresented in plasma fibrinogen and contributed only about 25% of the total Aα chain material.
Hereafter, all the studies performed to characterize the new abnormal fibrinogen were performed with the plasma of the proband’s father (indicated as patient), except the fibrin polymerisation and the fibrin clot growth assessment performed in both proband and proband’s father.

Fibrinogen degradation, western blotting and fibrin factor XIIIa cross-linking

Plasmin attacks the middle of the coiled-coil of fibrinogen and generates the degradation products fragments Y and D. Surprisingly, the degradation of patient fibrinogen (Aα Arg104 > Cys) by plasmin was similar to control in the presence of either Ca2+ or EDTA (results not shown). In addition, the mutation introduces an unpaired –SH that could potentially form fibrinogen-albumin complexes, although by immunoblotting fibrinogen binding to albumin was negative (results not shown). The patient fibrin α-chain factor XIIIa cross-linking seemed faster compared to control. It can be seen in Fig. 2 that the patient has more intense bands of higher molecular weight corresponding to α-chain factor XIIIa cross-linking at short incubation times (i.e. 2, 5, and 15 min) compared to control.

Fibrin polymerization and Fibrinolysis

Fibrin formation in the proband’s plasma had a near normal profile with only a slightly decreased final turbidity (~12%), while in the father the fibrin fibers growth (reflected in the slope value) and consequently the final turbidity was decreased approximately 1.3× and 40%, respectively (Table 2, Fig. 3a). With purified fibrinogen, the proband’s father polymerization was similar to control (Fig. 3b).
Table 2
Plasma fibrin polymerization. Plasma polymerization was done with fresh plasma. The optical density (OD) was multiplied by 1000 (mOD). The results are presented as mean (±SD)
Parameters
Control
Proband
Father
Mother
Fg* [mg/dl]
258
183
168
523
Lag time (s)
0
0
7.5 ± 11
0
Slope (mOD/s)
2.23 ± 0.19
1.66 ± 0.28
0.86 ± 0.24
2.38 ± 0.58
MaxAbs (mOD)
833 ± 18
736 ± 0.6
520 ± 4
970 ± 31
* Fg: fibrinogen, by clot weight method [11]
MaxAbs maximum absorbance
The father’s fibrin clot dissolution had a slightly shorter T50: 485 ± 33 s compared to 613 ± 62 s in control (p = 0.001), and a slightly delayed fibrinolysis rate: 1.9 ± 0.4 × 10−4 OD/s compared to 2.5 ± 0.4 × 10−4 OD/s in control (p = 0.008). In Fig. 4a are shown the fibrinolysis curves, the patient area under the curve (AUC) was 1.54 × 107 compared to 3.64 × 107 control, approximately 2× difference, and in Fig. 4b the T50 s and slopes distribution are represented.

Clots biophysical characterization

The patient fibrin elastic modulus was approximately 500 dyne/cm2 less than the control, but this was not statistical significant (Table 3). The patient clot’s surface available for flux (Ks) was almost 1.8× higher than control (p < <0.001). The confocal microscopy images showed subtle differences between the patient fibrin network and control (Fig. 5). The patient fibrin density and diameter were 0.329 ± 0.016 peaks/μm and 1.150 ± 0.642 μm, respectively, compared to 0.316 peaks/μm ± 0.017 and 1.23 ± 0.02 μm in the control (p < 0.05). In the orthogonal fibrin clot view the patient clot looked more porous, which correlated with the porosity value (Ks) that was approximately 2× higher.
Table 3
Summary of clots biophysical characterization. Results are showed as the mean (±SD). In brackets the number of values averaged
 
Control
Father
p
Fg (mg/dl)
222
176
 
EM (dyne/cm2)
3239 ± 671
(n = 9)
2734 ± 534
(n = 7)
0.116
Ks ×10−9
(cm2)
5.1 ± 0.9
(n = 20)
9.4 ± 1
(n = 22)
1.83 × 10−17
Fg: fibrinogen concentration by clot weight method. EM: elastic modulus. Ks: permeation constant

Thrombodynamics

As shown in Table 4, all the values of thrombodynamics parameters were decreased in the proband and her father compared to the healthy control. The lag time, the initial rate and the clot size at 30 min were similar between the proband and her father. The clot density was lower in father compared to the proband (9625.5 versus 11,206 arbitrary unit) reflecting his lower fibrinogen concentration. Videos of dynamic clots growth are shown in Additional file 1 (control), Additional file 2 (father) and Additional file 3 (proband).
Table 4
Thrombodynamics data
 
Control
Proband
Father
Lag time (min)
1.1
1.45
1.4
Initial rate (μm/min)
57.45
57.5
63.3
Stationary rate (μm/min)
56.4
32.1
33.7
Size at 30 min (μm)
1814
1306.5
1368.6
Density (a.u.)
22,161
11,026
9626.5
Min minutes, a.u. arbitrary unit

Discussion

A new fibrinogen mutation was found accidentally in an 11 year-old girl when routine pre-operative coagulation tests were performed. Gene sequencing revealed a missense mutation in FGA: g. 3057 C > T that predicts a p. Arg104 > Cys substitution. The low functional and antigenic fibrinogen concentrations in the father and daughter can be explained by the low expression level of the Aα Arg104 → Cys chain in their plasma fibrinogen as the new variant made up only about 25% of the total. Also, because the Aα Arg104 → Cys mutation creates a potential N-glycosylation site (Asn-Asn-Cys) centered on Asn103, mass specters were carefully examined for any new Aα chains with bi-antennary oligosaccharide side chains, but none were observed at or around their expected position at +2202 Da.
The fibrinogen coiled-coil connects the central E nodule with the distal nodule D and is made up of Aα chain residues 50–160, Bβ 81–191 and γ 24–134 [3]. At the middle of the coiled coil the structure become disorganized, probably due to the presence of proline residues at γ 70 and 76 [3]. There are 19 afibrinogenemia, 6 hypofibrinogenemia and 2 dysfibrinogenemia mutations reported in the stretch of Aα 100–112 (http://​www.​geht.​org). At present, there are only 3 hypodysfibrinogenemia reported in the middle of the coiled coil region: Fibrinogen Epsom (Bβ Asn137_Glu141), Michigan (γ Tyr114His), and Leipzig II (Ala82Gly) [6], although the last two are compound heterozygous. These naturally occurring fibrinogen variants confirm previous findings about the role of the helical coiled coil in fibrinogen biosynthesis and secretion [10]. Other mutations in this region cause dysfibrinogenemia. For example, fibrinogen Plzen (Aα Asn106Asp) and Vizovice (Aα Phe98Ile) are characterized by low functional fibrinogen, 1.13 g/l and 1.66, but normal immunologic fibrinogen concentration 3.99 g/l and 2.89, respectively [23, 24]. The normal fibrinogen antigenic concentration excludes that these abnormal fibrinogens have impaired secretion. The low fibrinogen activity highlight the role of this part of the molecule on fibrinogen polymerization.
The kinetics of fibrin formation in plasma studied by turbidity was more impaired in the father compared to the proband. However, by thrombodynamics the father and proband behaved similarly. The results obtained using this last technique seemed more consistent with mass spectrometry. Similar polymerization kinetic would be expected for both individuals as difference of 0.15 g/l in their functional fibrinogen concentrations appeared insufficient to cause major changes polymerization. These discrepant results could result from the fact that the thrombodynamics uses a primarily different principle to initiate the coagulation cascade compared to the turbidity assay. In the former, the clotting is activated by a surface with immobilized tissue factor, which better resembles in vivo clot formation, whereas in the latter the clotting is activated by homogeneously dissolved thrombin. Fibrinogen functional correlations between these two assays require further studies.
We also expected the father would exhibit altered fibrin formation in a purified system, but probably the mutated molecules were underrepresented compared to plasma. In addition, we do not know if the mutated molecules were lost during the purification process. In the future it would be advisable for these cases to choose an immunopurification method.
The proband had higher fibrinogen concentration than her father, despite being about 5 decades younger. Different studies have found that fibrinogen increases with age [2527]. Hager et al. reported a 25 mg/dl increases per decade [26]. Since fibrinogen is an acute phase reactant protein and CRP was not measured, it cannot be ruled out that the proband at the time of blood withdrawal could have an infectious disease.
When analyzed, an hypofibrinolysis has been observed in most hypodysfibrinogenemia [6]. The Aα Arg104 and Bβ Lys133 are the very first plasmin attack points. However, the fibrinolytic process of fibrinogen Caracas IX was close to normal. In contrast, fibrinogen Epsom with a deletion of the residues Asn137_Glu141 showed hyperfibrinolysis, together with an increased fibrinogen clearance [28]. In fibrinogen Dunedin whose γ82Ala → Gly mutation occurs near the plasmin sensitive site in the coiled coil region also displayed an increased proteolytic sensitivity. It would be interesting in the future to explore the causes of these differences in order to shed light on possibly unknown mechanisms of fibrin(o)gen degradation.

Conclusions

The fibrinogen mutation Aα Arg104 > Cys did not introduce relevant clinical consequences probably due to its low expression levels.

Acknowledgements

We are grateful to the patients of the present study for their collaboration.

Funding

The investigation has been partially supported by the government of Bolivarian Republic of Venezuela.

Availability of data and materials

Most data generated or analyzed during this study are included in this published article (and its supplementary information files). Other datasets used and/or analyzed during the current study and not shown are available from the corresponding author on reasonable request.
The Ethics Committee of the Instituto Venezolano de Investigaciones Científicas (IVIC) approved the project under the number DIR-1925/1579/2016. All subjects signed an informed consent before blood withdrawal, and written parental consent was obtained for the proband.
Not applicable.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided 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 Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated.
Literatur
1.
Zurück zum Zitat Blomback B, Hessel B, Hogg D, Therkildsen L. A two-step fibrinogen--fibrin transition in blood coagulation. Nature. 1978;275:501–5.CrossRefPubMed Blomback B, Hessel B, Hogg D, Therkildsen L. A two-step fibrinogen--fibrin transition in blood coagulation. Nature. 1978;275:501–5.CrossRefPubMed
2.
Zurück zum Zitat Asselta R, Duga S, Tenchini ML. The molecular basis of quantitative fibrinogen disorders. J Thromb Haemost. 2006;4:2115–29.CrossRefPubMed Asselta R, Duga S, Tenchini ML. The molecular basis of quantitative fibrinogen disorders. J Thromb Haemost. 2006;4:2115–29.CrossRefPubMed
3.
Zurück zum Zitat Doolittle RF, Goldbaum DM, Doolittle LR. Designation of sequences involved in the “coiled-coil” interdomainal connections in fibrinogen: constructions of an atomic scale model. J Mol Biol. 1978;120:311–25.CrossRefPubMed Doolittle RF, Goldbaum DM, Doolittle LR. Designation of sequences involved in the “coiled-coil” interdomainal connections in fibrinogen: constructions of an atomic scale model. J Mol Biol. 1978;120:311–25.CrossRefPubMed
4.
Zurück zum Zitat Neerman-Arbez M, de Moerloose P, Casini A. Laboratory and genetic investigation of mutations accounting for congenital fibrinogen disorders. Semin Thromb Hemost. 2016;42:356–65.CrossRefPubMed Neerman-Arbez M, de Moerloose P, Casini A. Laboratory and genetic investigation of mutations accounting for congenital fibrinogen disorders. Semin Thromb Hemost. 2016;42:356–65.CrossRefPubMed
5.
Zurück zum Zitat Cunningham MT, Brandt JT, Laposata M, Olson JD. Laboratory diagnosis of dysfibrinogenemia. Arch Pathol Lab Med. 2002;126:499–505.PubMed Cunningham MT, Brandt JT, Laposata M, Olson JD. Laboratory diagnosis of dysfibrinogenemia. Arch Pathol Lab Med. 2002;126:499–505.PubMed
6.
Zurück zum Zitat Casini A, Brungs T, Lavenu-Bombled C, Vilar R, Neerman-Arbez M, de Moerloose P. Genetics, diagnosis and clinical features of congenital hypodysfibrinogenemia: a systematic literature review and report of a novel mutation. J Thromb Haemost. 2017;15:876–88 Casini A, Brungs T, Lavenu-Bombled C, Vilar R, Neerman-Arbez M, de Moerloose P. Genetics, diagnosis and clinical features of congenital hypodysfibrinogenemia: a systematic literature review and report of a novel mutation. J Thromb Haemost. 2017;15:876–88
7.
Zurück zum Zitat Maghzal GJ, Brennan SO, Homer VM, George PM. The molecular mechanisms of congenital hypofibrinogenaemia. Cell Mol Life Sci. 2004;61:1427–38.CrossRefPubMed Maghzal GJ, Brennan SO, Homer VM, George PM. The molecular mechanisms of congenital hypofibrinogenaemia. Cell Mol Life Sci. 2004;61:1427–38.CrossRefPubMed
8.
Zurück zum Zitat Zhang JZ, Redman CM. Role of interchain disulfide bonds on the assembly and secretion of human fibrinogen. J Biol Chem. 1994;269:652–8.PubMed Zhang JZ, Redman CM. Role of interchain disulfide bonds on the assembly and secretion of human fibrinogen. J Biol Chem. 1994;269:652–8.PubMed
9.
Zurück zum Zitat Zhang JZ, Redman C. Fibrinogen assembly and secretion. Role of intrachain disulfide loops. J Biol Chem. 1996;271:30083–8.CrossRefPubMed Zhang JZ, Redman C. Fibrinogen assembly and secretion. Role of intrachain disulfide loops. J Biol Chem. 1996;271:30083–8.CrossRefPubMed
10.
Zurück zum Zitat Xu W, Chung DW, Davie EW. The assembly of human fibrinogen. The role of the amino-terminal and coiled-coil regions of the three chains in the formation of the alphagamma and betagamma heterodimers and alphabetagamma half-molecules. J Biol Chem. 1996;271:27948–53.CrossRefPubMed Xu W, Chung DW, Davie EW. The assembly of human fibrinogen. The role of the amino-terminal and coiled-coil regions of the three chains in the formation of the alphagamma and betagamma heterodimers and alphabetagamma half-molecules. J Biol Chem. 1996;271:27948–53.CrossRefPubMed
12.
Zurück zum Zitat Niwa K, Takebe M, Sugo T, Kawata Y, Mimuro J, Asakura S, et al. A gamma Gly-268 to Glu substitution is responsible for impaired fibrin assembly in a homozygous dysfibrinogen Kurashiki I. Blood. 1996;87:4686–94.PubMed Niwa K, Takebe M, Sugo T, Kawata Y, Mimuro J, Asakura S, et al. A gamma Gly-268 to Glu substitution is responsible for impaired fibrin assembly in a homozygous dysfibrinogen Kurashiki I. Blood. 1996;87:4686–94.PubMed
13.
Zurück zum Zitat Marchi R, Walton BL, McGary CS, Lin FC, Ma AD, Pawlinski R, et al. Dysregulated coagulation associated with hypofibrinogenaemia and plasma hypercoagulability: implications for identifying coagulopathic mechanisms in humans. Thromb Haemost. 2012;108:516–26.CrossRefPubMedPubMedCentral Marchi R, Walton BL, McGary CS, Lin FC, Ma AD, Pawlinski R, et al. Dysregulated coagulation associated with hypofibrinogenaemia and plasma hypercoagulability: implications for identifying coagulopathic mechanisms in humans. Thromb Haemost. 2012;108:516–26.CrossRefPubMedPubMedCentral
14.
Zurück zum Zitat Marchi R, Carvajal Z, Meyer M, Soria J, Ruiz-Saez A, Arocha-Pinango CL, et al. Fibrinogen Guarenas, an abnormal fibrinogen with an Aalpha-chain truncation due to a nonsense mutation at Aalpha 467 Glu (GAA)-->stop (TAA). Thromb Res. 2006;118:637–50.CrossRefPubMed Marchi R, Carvajal Z, Meyer M, Soria J, Ruiz-Saez A, Arocha-Pinango CL, et al. Fibrinogen Guarenas, an abnormal fibrinogen with an Aalpha-chain truncation due to a nonsense mutation at Aalpha 467 Glu (GAA)-->stop (TAA). Thromb Res. 2006;118:637–50.CrossRefPubMed
15.
Zurück zum Zitat Towbin H, Staehelin T, Gordon J. Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proc Natl Acad Sci U S A. 1979;76:4350–4.CrossRefPubMedPubMedCentral Towbin H, Staehelin T, Gordon J. Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proc Natl Acad Sci U S A. 1979;76:4350–4.CrossRefPubMedPubMedCentral
16.
Zurück zum Zitat Marchi R, Echenagucia M, Meyer M, Acosta M, Apitz R, Ruiz-Sáez A. Fibrinogen Maracaibo: Hypo-Dysfibrinogenemia Caused by a Heterozygous Mutation in the Gen that Encodes for the Fibrinogen Aα Chain (G.1194G>A: P.Gly13>Glu) with Diminished Thrombin Generation. J Blood Disorders Transf. 2014;5:1–5. Marchi R, Echenagucia M, Meyer M, Acosta M, Apitz R, Ruiz-Sáez A. Fibrinogen Maracaibo: Hypo-Dysfibrinogenemia Caused by a Heterozygous Mutation in the Gen that Encodes for the Fibrinogen Aα Chain (G.1194G>A: P.Gly13>Glu) with Diminished Thrombin Generation. J Blood Disorders Transf. 2014;5:1–5.
17.
Zurück zum Zitat Carter AM, Cymbalista CM, Spector TD, Grant PJ. Heritability of clot formation, morphology, and lysis: the EuroCLOT study. Arterioscler Thromb Vasc Biol. 2007;27:2783–9.CrossRefPubMed Carter AM, Cymbalista CM, Spector TD, Grant PJ. Heritability of clot formation, morphology, and lysis: the EuroCLOT study. Arterioscler Thromb Vasc Biol. 2007;27:2783–9.CrossRefPubMed
18.
Zurück zum Zitat Brennan SO, Mangos H, Faed JM. Benign FGB (148Lys-->Asn, and 448Arg-->Lys), and novel causative gamma211Tyr-->his mutation distinguished by time of flight mass spectrometry in a family with hypofibrinogenaemia. Thromb Haemost. 2014;111:679–84.CrossRefPubMed Brennan SO, Mangos H, Faed JM. Benign FGB (148Lys-->Asn, and 448Arg-->Lys), and novel causative gamma211Tyr-->his mutation distinguished by time of flight mass spectrometry in a family with hypofibrinogenaemia. Thromb Haemost. 2014;111:679–84.CrossRefPubMed
19.
Zurück zum Zitat Pieters M, Undas A, Marchi R, De Maat MP, Weisel J, Ariens RA. An international study on the standardization of fibrin clot permeability measurement: methodological considerations and implications for healthy control values. J Thromb Haemost. 2012;10:2179–81.CrossRefPubMed Pieters M, Undas A, Marchi R, De Maat MP, Weisel J, Ariens RA. An international study on the standardization of fibrin clot permeability measurement: methodological considerations and implications for healthy control values. J Thromb Haemost. 2012;10:2179–81.CrossRefPubMed
20.
21.
Zurück zum Zitat Potze W, Adelmeijer J, Porte RJ, Lisman T. Preserved clot formation detected by the Thrombodynamics analyzer in patients with cirrhosis. Thromb Res. 2015;135:1012–6.CrossRefPubMed Potze W, Adelmeijer J, Porte RJ, Lisman T. Preserved clot formation detected by the Thrombodynamics analyzer in patients with cirrhosis. Thromb Res. 2015;135:1012–6.CrossRefPubMed
22.
Zurück zum Zitat Panteleev MA, Ovanesov MV, Kireev DA, Shibeko AM, Sinauridze EI, Ananyeva NM, et al. Spatial propagation and localization of blood coagulation are regulated by intrinsic and protein C pathways, respectively. Biophys J. 2006;90:1489–500.CrossRefPubMed Panteleev MA, Ovanesov MV, Kireev DA, Shibeko AM, Sinauridze EI, Ananyeva NM, et al. Spatial propagation and localization of blood coagulation are regulated by intrinsic and protein C pathways, respectively. Biophys J. 2006;90:1489–500.CrossRefPubMed
23.
Zurück zum Zitat Riedelova-Reicheltova Z, Kotlin R, Suttnar J, Geierova V, Riedel T, Majek P, et al. A novel natural mutation AalphaPhe98Ile in the fibrinogen coiled-coil affects fibrinogen function. Thromb Haemost. 2014;111:79–87.CrossRefPubMed Riedelova-Reicheltova Z, Kotlin R, Suttnar J, Geierova V, Riedel T, Majek P, et al. A novel natural mutation AalphaPhe98Ile in the fibrinogen coiled-coil affects fibrinogen function. Thromb Haemost. 2014;111:79–87.CrossRefPubMed
24.
Zurück zum Zitat Kotlin R, Reicheltova Z, Maly M, Suttnar J, Sobotkova A, Salaj P, et al. Two cases of congenital dysfibrinogenemia associated with thrombosis - fibrinogen Praha III and fibrinogen Plzen. Thromb Haemost. 2009;102:479–86.PubMed Kotlin R, Reicheltova Z, Maly M, Suttnar J, Sobotkova A, Salaj P, et al. Two cases of congenital dysfibrinogenemia associated with thrombosis - fibrinogen Praha III and fibrinogen Plzen. Thromb Haemost. 2009;102:479–86.PubMed
25.
Zurück zum Zitat Krobot K, Hense HW, Cremer P, Eberle E, Keil U. Determinants of plasma fibrinogen: relation to body weight, waist-to-hip ratio, smoking, alcohol, age, and sex. Results from the second MONICA Augsburg survey 1989–1990. Arterioscler Thromb. 1992;12:780–8.CrossRefPubMed Krobot K, Hense HW, Cremer P, Eberle E, Keil U. Determinants of plasma fibrinogen: relation to body weight, waist-to-hip ratio, smoking, alcohol, age, and sex. Results from the second MONICA Augsburg survey 1989–1990. Arterioscler Thromb. 1992;12:780–8.CrossRefPubMed
26.
Zurück zum Zitat Hager K, Felicetti M, Seefried G, Platt D. Fibrinogen and aging. Aging (Milano). 1994;6:133–8. Hager K, Felicetti M, Seefried G, Platt D. Fibrinogen and aging. Aging (Milano). 1994;6:133–8.
27.
Zurück zum Zitat Spada RS, Toscano G, Chiarenza S, Di Mauro S, Cosentino FI, Iero I, et al. Ischemic stroke and fibrinogen in the elderly. Arch Gerontol Geriatr Suppl. 2004;9:403–6.CrossRef Spada RS, Toscano G, Chiarenza S, Di Mauro S, Cosentino FI, Iero I, et al. Ischemic stroke and fibrinogen in the elderly. Arch Gerontol Geriatr Suppl. 2004;9:403–6.CrossRef
28.
Zurück zum Zitat Brennan SO, Davis RL, Lowen R, Ruskova A. Deletion of five residues from the coiled coil of fibrinogen (Bbeta Asn167_Glu171del) associated with bleeding and hypodysfibrinogenemia. Haematologica. 2009;94:585–8.CrossRefPubMedPubMedCentral Brennan SO, Davis RL, Lowen R, Ruskova A. Deletion of five residues from the coiled coil of fibrinogen (Bbeta Asn167_Glu171del) associated with bleeding and hypodysfibrinogenemia. Haematologica. 2009;94:585–8.CrossRefPubMedPubMedCentral
Metadaten
Titel
A novel fibrinogen mutation: FGA g. 3057 C > T (p. Arg104 > Cys) impairs fibrinogen secretion
verfasst von
R. Marchi
M. Linares
H. Rojas
A. Ruiz-Sáez
M. Meyer
A. Casini
S.O. Brennan
Publikationsdatum
01.12.2017
Verlag
BioMed Central
Erschienen in
BMC Hematology / Ausgabe 1/2017
Elektronische ISSN: 2052-1839
DOI
https://doi.org/10.1186/s12878-017-0086-8

Weitere Artikel der Ausgabe 1/2017

BMC Hematology 1/2017 Zur Ausgabe

Adjuvante Immuntherapie verlängert Leben bei RCC

25.04.2024 Nierenkarzinom Nachrichten

Nun gibt es auch Resultate zum Gesamtüberleben: Eine adjuvante Pembrolizumab-Therapie konnte in einer Phase-3-Studie das Leben von Menschen mit Nierenzellkarzinom deutlich verlängern. Die Sterberate war im Vergleich zu Placebo um 38% geringer.

Alectinib verbessert krankheitsfreies Überleben bei ALK-positivem NSCLC

25.04.2024 NSCLC Nachrichten

Das Risiko für Rezidiv oder Tod von Patienten und Patientinnen mit reseziertem ALK-positivem NSCLC ist unter einer adjuvanten Therapie mit dem Tyrosinkinase-Inhibitor Alectinib signifikant geringer als unter platinbasierter Chemotherapie.

Bei Senioren mit Prostatakarzinom auf Anämie achten!

24.04.2024 DGIM 2024 Nachrichten

Patienten, die zur Behandlung ihres Prostatakarzinoms eine Androgendeprivationstherapie erhalten, entwickeln nicht selten eine Anämie. Wer ältere Patienten internistisch mitbetreut, sollte auf diese Nebenwirkung achten.

ICI-Therapie in der Schwangerschaft wird gut toleriert

Müssen sich Schwangere einer Krebstherapie unterziehen, rufen Immuncheckpointinhibitoren offenbar nicht mehr unerwünschte Wirkungen hervor als andere Mittel gegen Krebs.

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.