Luteinizing hormone (LH) is a glycoproteingonadotropin hormone secreted by the gonadotroph cells of the anterior pituitary gland, working in tandem with follicle-stimulating hormone (FSH) to regulate reproductive functions in both sexes.[1] Produced in response to pulsatile gonadotropin-releasing hormone (GnRH) from the hypothalamus, LH levels are modulated by negative feedback from sex steroids—estrogen and progesterone in females, and testosterone in males—ensuring coordinated gonadal activity.[1] Essential for puberty, sexual maturation, and fertility, LH drives key reproductive processes, including ovulation in females and spermatogenesis support in males.[2]In females, LH plays a pivotal role across the menstrual cycle: during the follicular phase, it collaborates with FSH to promote ovarian follicle development and estrogen production; a mid-cycle surge in LH, triggered by rising estrogen levels, induces ovulation by causing the mature follicle to rupture and release the oocyte.[1] Post-ovulation, LH sustains the corpus luteum, facilitating progesterone secretion to prepare the uterus for potential implantation.[1] Dysregulation of LH, such as in polycystic ovary syndrome (PCOS) where elevated levels contribute to hyperandrogenism,[3] or in hypogonadotropic hypogonadism with deficient secretion leading to infertility,[4] underscores its clinical importance.In males, LH binds to receptors on Leydig cells in the testes, stimulating the synthesis and release of testosterone, which is vital for spermatogenesis, secondary sexual characteristics, and libido maintenance.[1] Throughout life, LH secretion remains relatively stable in a pulsatile pattern, contrasting the cyclic fluctuations in females, and its measurement via blood tests helps diagnose conditions like Klinefelter syndrome (high LH with low testosterone) or androgen deficiency.[5] Overall, LH exemplifies the intricate hypothalamic-pituitary-gonadal axis, integrating neural and endocrine signals to orchestrate reproduction.[1]
Discovery and Nomenclature
Etymology
The term "luteinizing hormone" (LH) originates from its primary function in promoting luteinization, the transformation of the ruptured ovarian follicle into the corpus luteum following ovulation in females.[6] This naming reflects the hormone's key role in reproductive physiology, where it surges to trigger ovulation and subsequent corpus luteum formation to support early pregnancy.[7]The linguistic root of "luteinizing" traces to "lutein," derived from the Latin lūteum (egg yolk) or lūteus (yellow), alluding to the yellowish appearance of the corpus luteum due to its lipid-rich cells.[8] The term corpus luteum itself combines Latin corpus (body) and lūteum (yellow), coined in the 19th century to describe this endocrine structure.[9]In the context of gonadotropins, LH was historically distinguished from follicle-stimulating hormone (FSH) through early 20th-century nomenclature; initially termed "Prolan B" for its luteinizing effects, it was later renamed LH to emphasize its specific induction of the corpus luteum, while FSH (formerly "Prolan A") was named for follicular development.[10] This convention underscores the complementary actions of these pituitary gonadotropins in regulating gonadal function.[11]
Historical Discovery
The initial recognition of gonadotropic activity in the pituitary gland emerged in the mid-1920s through experiments demonstrating its essential role in gonadal function. In 1926, American endocrinologist Philip E. Smith reported that surgical removal of the pituitary in rats caused rapid atrophy of the ovaries and testes, but implanting a functional pituitary from a donor animal restored gonadal development, providing early evidence for pituitary-derived gonad-stimulating substances. Building on this, Smith and his collaborator Earl T. Engle extracted active gonadotropic material from bovine anterior pituitary glands in 1927, showing it could induce follicular growth and luteinization in hypophysectomized rats when administered.[12]Concurrently, German researcher Bernhard Zondek advanced the field by investigating pituitary hormones in 1926, demonstrating that extracts from the anterior pituitary induced estrus and ovulation in immature female rats, highlighting their gonadotropic potency and contributing to foundational insights that influenced later endocrine research, including Nobel Prize considerations for pituitary hormone studies.[13] By the early 1930s, efforts to differentiate the gonadotropins intensified using bioassays based on specific ovarian and testicular responses. In 1931, Howard L. Fevold, F. L. Hisaw, and Sidney L. Leonard separated pituitary extracts into two distinct fractions: one promoting follicular development (later identified as follicle-stimulating hormone, FSH) and another inducing luteinization and interstitial cell activity (luteinizing hormone, LH), confirmed through assays in rats and rabbits.[14]Purification techniques progressed significantly in the 1940s, enabling more precise characterization. In 1940, biochemist Choh Hao Li achieved the first isolation of LH from sheep anterior pituitary glands using acid-acetone extraction and precipitation methods, yielding a highly active preparation that specifically stimulated luteinization without substantial FSH contamination.[15] These advances facilitated early clinical applications in the 1940s and 1950s, as purified animal-derived gonadotropins were tested for infertility treatment; by 1950, human menopausal gonadotropin (hMG), rich in both FSH and LH and extracted from the urine of postmenopausal women, was introduced for ovulation induction in anovulatory women, advancing human-derived gonadotropin-based fertility therapies.[10]
Molecular Biology
Gene Structure and Expression
The luteinizing hormone beta subunit gene, LHB, is located on the long arm of human chromosome 19 at position 19q13.3, spanning approximately 1.1 kb of genomic DNA.[16] This gene encodes the precursor for the beta subunit of luteinizing hormone (LH), a glycoprotein hormone essential for reproduction. The LHB gene consists of three exons separated by two introns, with exon 1 encoding the signal peptide and part of the mature protein, exon 2 containing the majority of the coding sequence, and exon 3 completing the mature beta subunit along with the 3' untranslated region.[17] The promoter region upstream of exon 1 includes responsive elements to gonadotropin-releasing hormone (GnRH), such as EGR1-binding sites, which mediate transcriptional activation in response to pulsatile GnRH stimulation from the hypothalamus.[18]Expression of LHB is highly tissue-specific, occurring predominantly in the gonadotroph cells of the anterior pituitary gland, where it is co-expressed with the alpha subunit gene (CGA) to form the functional LH heterodimer.[19] This restricted expression pattern ensures that LH production is confined to the pituitary, with levels tightly regulated by hypothalamic signals and feedback mechanisms. In these gonadotrophs, LHB transcription is dynamically controlled by GnRH pulses, which induce immediate early genes like Egr1 to bind the promoter and drive expression, adapting to physiological needs such as puberty or the menstrual cycle.[18]Mutations and polymorphisms in LHB are associated with reproductive disorders, particularly inactivating mutations that lead to selective luteinizing hormone deficiency and hypogonadotropic hypogonadism. For instance, homozygous nonsense or missense mutations, such as Trp28X or those altering exon 2 sequences, disrupt beta subunit production, resulting in low or undetectable LH levels, impaired gonadal function, infertility, and delayed puberty in affected individuals.[20][21] These variants highlight the gene's critical role, as even rare homozygous changes can cause profound endocrine disruption due to the absence of functional LH.The LHB gene exhibits strong evolutionary conservation across mammals, with orthologs identified in at least 28 species ranging from rodents to primates, reflecting its fundamental role in reproductive physiology.[22] This conservation extends to the exon-intron structure and key regulatory elements, such as GnRH-responsive promoters, which are preserved from rats to humans, underscoring the gene's ancient origin within the glycoprotein hormone family.
Protein Structure and Subunits
Luteinizing hormone (LH) is a heterodimeric glycoprotein composed of two noncovalently linked subunits, α and β, which together confer its biological specificity and activity. The α subunit, common to all pituitary glycoprotein hormones including follicle-stimulating hormone (FSH), thyroid-stimulating hormone (TSH), and chorionic gonadotropin (CG), consists of 92 amino acids and is encoded by the CGA gene on chromosome 6q14.3. This subunit features two N-linked glycosylation sites at asparagine residues 52 and 78, which contribute to the hormone's stability and half-life. The β subunit, unique to LH, comprises 121 amino acids and contains two N-linked glycosylation sites at asparagines 13 and 30; these sites are critical for proper folding, secretion, and bioactivity, as mutations or alterations here can lead to aggregation or reduced function.[7][23][24]The quaternary structure of LH arises from the noncovalent association of the α and β subunits, forming a compact heterodimer. The alpha subunit is stabilized by ten conserved cysteine residues that form five disulfide bridges, while the beta subunit has twelve cysteines forming six disulfide bridges; both contribute to the characteristic "cysteine knot" motif, essential for the hormone's three-dimensional conformation and receptor binding. Glycosylation, primarily N-linked, adds complex carbohydrate chains that account for approximately 20-30% of the molecule's mass, enhancing solubility, protecting against proteolysis, and modulating clearance rates. The resulting molecular weight of intact human LH is about 28-30 kDa, with the carbohydrate component varying slightly based on sialic acid content and isoform heterogeneity.[25][26]In comparison to human chorionic gonadotropin (hCG), the LH β subunit lacks the extended C-terminal peptide (approximately 24 amino acids) present in hCG β, which bears additional O-linked glycosylation sites and prolongs hCG's circulatory half-life. This structural difference underscores LH's shorter duration of action, typically around 20-60 minutes, suited to its role in acute gonadal stimulation. Overall, the subunit architecture and post-translational modifications of LH exemplify the evolutionary conservation and functional diversification within the glycoprotein hormone family.[27]
Biosynthesis and Regulation
Synthesis in the Anterior Pituitary
Luteinizing hormone (LH) is synthesized by specialized gonadotroph cells, also known as gonadotrope cells, within the anterior pituitary gland. These cells belong to the gonadotrope lineage and are responsible for producing both LH and follicle-stimulating hormone (FSH), sharing a common alpha subunit but differing in their beta subunits. The synthesis process begins with the transcription of the LH beta subunit gene (LHB) in the nucleus of these gonadotrophs, driven by pituitary-specific transcription factors such as steroidogenic factor-1 (SF-1, also known as NR5A1) and LIM homeodomain transcription factor 3 (LHX3). SF-1 binds to specific promoter regions of the LHB gene to enhance its expression, while LHX3 contributes to cell-specific activation in gonadotrophs. Additional factors like GATA2 and activin signaling pathways further modulate LHB transcription, ensuring regulated production in response to physiological needs.[28][29][30][31]Following transcription, the LH subunits undergo processing in the endoplasmic reticulum (ER), where disulfide bond formation stabilizes their folding through the action of protein disulfide isomerases. The nascent polypeptides are then transported to the Golgi apparatus for post-translational modifications, primarily N-linked glycosylation, which adds complex carbohydrate chains to asparagine residues on both the alpha and beta subunits. These glycosylation events are crucial for proper subunit assembly into the heterodimeric LH molecule, influencing its biological activity, half-life, and secretion efficiency. The fully processed LH is packaged into secretory granules within the gonadotrophs for storage.[25][32][7]Mature LH is stored in these secretory granules and released in a pulsatile manner from the anterior pituitary, synchronized with physiological rhythms to maintain reproductive function. This pulsatile release pattern arises from the intermittent stimulation by hypothalamic signals, allowing for dynamic control of LH output. Additionally, synthesis rates are subject to negative feedback inhibition by gonadal steroids, such as estrogen and testosterone, which act directly on gonadotrophs to suppress LHB gene transcription and reduce overall production. For instance, estrogen binds to nuclear receptors in these cells, repressing promoter activity and thereby fine-tuning LH levels to prevent overproduction.[33][34][35][36]
Hypothalamic Regulation
The secretion of luteinizing hormone (LH) from the anterior pituitary is primarily regulated by gonadotropin-releasing hormone (GnRH), which is produced and released by specialized neurons in the preoptic area and arcuate nucleus of the hypothalamus.[37] GnRH is secreted into the hypophyseal portal circulation in a pulsatile manner, with pulses occurring approximately every 1-2 hours depending on the reproductive phase, directly stimulating gonadotroph cells in the pituitary to synthesize and release LH in episodic bursts that mirror the hypothalamic rhythm.[37] This pulsatile pattern is essential for maintaining normal gonadotropin secretion; continuous GnRH exposure leads to desensitization and downregulation of pituitary GnRH receptors, suppressing LH release.[38]The frequency and amplitude of GnRH pulses influence the relative secretion of LH and follicle-stimulating hormone (FSH), with higher pulse frequencies favoring LH production over FSH due to differential gene expression in gonadotrophs.[38] In females, sex steroids exert negative feedback on GnRH secretion: estradiol and progesterone from the ovaries inhibit GnRH pulse frequency and amplitude, thereby reducing LH levels to prevent premature ovulation during the follicular phase.[39] In males, testosterone provides negative feedback by suppressing hypothalamic GnRH release, maintaining steady-state LH secretion that supports spermatogenesis without surges.[39]During the preovulatory phase in females, prolonged estrogen exposure triggers a switch to positive feedback, priming the hypothalamus to increase GnRH pulse frequency and amplitude, culminating in a massive LH surge that induces ovulation.[40] This ovulatory surge is facilitated by estrogen-sensitive neurons in the anteroventral periventricular nucleus, which enhance GnRH neuron excitability.[37] Upstream modulators, including kisspeptin and neurokinin B (NKB) expressed in KNDy neurons (co-expressing kisspeptin, NKB, and dynorphin) within the arcuate nucleus, form the core of the GnRH pulse generator; kisspeptin directly stimulates GnRH neurons via the KISS1R receptor to drive pulsatile release, while NKB autoregulates KNDy neuron activity to set pulse frequency, and dynorphin provides inhibitory tone.[41] Disruptions in this network alter LH pulsatility, underscoring its role in reproductive timing.[42]
Physiological Functions
Functions in Females
In females, luteinizing hormone (LH) is essential for the onset of puberty, where rising levels of LH, in coordination with follicle-stimulating hormone (FSH), stimulate the development of secondary sexual characteristics and initiate menarche, marking the first menstrual period.[43] During puberty, pulsatile LH secretion from the anterior pituitary increases, driving ovarian follicle maturation and estrogen production, which contribute to breast development, pubic hair growth, and the establishment of cyclic reproductive function.[44]Throughout the menstrual cycle, LH regulates ovarian activity in synergy with FSH to support follicular development. In the early follicular phase, LH binds to receptors on theca cells in developing ovarian follicles, stimulating these cells to produce androgens such as androstenedione, which serve as precursors for estrogen synthesis by granulosa cells under FSH influence.[45][46] This two-cell, two-gonadotropin model ensures coordinated steroidogenesis, with LH providing the androgen substrate that FSH-dependent aromatase enzymes convert to estradiol, promoting follicular growth and endometrial preparation.[47] LH levels remain relatively low during this phase but are critical for maintaining thecal cell proliferation and vascularization of the follicle.The mid-cycle LH surge, triggered by rising estradiol levels, is a pivotal event that induces ovulation by causing follicular rupture and oocyte release.[48] This surge, lasting approximately 24-48 hours, activates luteinizing hormone receptors on granulosa and theca cells, leading to increased expression of genes involved in proteolytic enzyme production, such as matrix metalloproteinases, which degrade the follicular wall and facilitate expulsion of the mature oocyte into the fallopian tube.[49] Post-ovulation, the LH surge promotes luteinization of the ruptured follicle, transforming granulosa and theca cells into the corpus luteum.In the luteal phase, LH sustains corpus luteum function by stimulating progesterone secretion, which is vital for endometrial thickening and preparation for potential implantation.[50] LH acts directly on luteal cells via cyclic AMP-mediated pathways to enhance cholesterol uptake and steroidogenic enzyme activity, ensuring peak progesterone output around days 21-23 of the cycle.[51] If pregnancy does not occur, declining LH levels contribute to corpus luteum regression, reducing progesterone and triggering menstruation, thus regulating the cycle's length and sequence.[52] Overall, LH's fluctuating concentrations orchestrate the menstrual cycle's progression from follicular recruitment to luteal support.[7]
Functions in Males
In males, luteinizing hormone (LH) primarily binds to receptors on the surface of Leydig cells in the testes, activating adenylate cyclase to increase intracellular cyclic AMP levels, which promotes the translocation of cholesterol into mitochondria for side-chain cleavage by the enzyme CYP11A1, initiating the biosynthesis of testosterone from pregnenolone.[53] This process is essential for steroidogenesis, as LH stimulation enhances the expression and activity of steroidogenic acute regulatory protein (StAR), facilitating cholesterol transport and subsequent conversion through enzymatic steps to produce testosterone.[54]Testosterone produced under LH influence diffuses locally to support spermatogenesis indirectly by acting on Sertoli cells and germ cells within the seminiferous tubules, promoting the maturation of spermatids into spermatozoa and maintaining the structural integrity of the blood-testis barrier.[55] Although follicle-stimulating hormone (FSH) directly targets Sertoli cells, LH-driven androgen production provides the high intratesticular testosterone concentrations necessary for completing meiosis and spermiogenesis, ensuring male fertility.[56]During puberty, the maturation of the hypothalamic-pituitary-gonadal axis leads to increasing pulsatile LH secretion, particularly at night, which drives the proliferation and differentiation of Leydig cells, resulting in testicular enlargement and elevated testosterone levels that induce virilization, including growth of secondary sexual characteristics such as facial and body hair, deepening of the voice, and penile development.[57] These nocturnal LH pulses herald the onset of puberty, correlating with initial signs of androgen action and progression through Tanner stages.[58]In adults, LH maintains Leydig cell function by sustaining testosterone production at levels required for reproductive health, with negative feedback loops where elevated testosterone and its metabolite estradiol inhibit LH secretion from the anterior pituitary via actions on gonadotropin-releasing hormone (GnRH) neurons and pituitary gonadotrophs, preventing overproduction.[59] Inhibin B, secreted by Sertoli cells in response to FSH, primarily modulates FSH feedback but contributes indirectly to the overall gonadal-pituitary balance influencing LH dynamics.[60]Through its mediation of androgen synthesis, LH supports broader physiological effects, including the maintenance of bone mineral density by promoting osteoblast activity and inhibiting osteoclast resorption in the skeletal system, as well as enhancing muscle mass and strength via androgen receptor signaling that stimulates protein synthesis and satellite cell activation in skeletal muscle.[61][62]
Functions in the Brain
Luteinizing hormone (LH) and its receptor, LHCGR, are expressed in various regions of the central nervous system, extending beyond their traditional roles in the reproductive axis. Studies utilizing single-cell transcriptomics have identified Lhcgr expression in over 400 brain regions, subregions, and nuclei across the mammalian brain, including significant levels in the hippocampus and cerebral cortex.[63] LHCGR is particularly dense in hippocampal and cortical neurons, where it facilitates direct signaling by LH that can cross the blood-brain barrier.[64] Additionally, LH itself is synthesized locally in the brain.[65]Emerging evidence from animal models highlights LH's neuroprotective functions, particularly in promoting neuronal survival and synaptic plasticity. Activation of CNS LHCGR has been shown to enhance cognition in ovariectomized female mice, an effect mediated through downstream signaling pathways like ERK, which supports learning, memory, and synaptic plasticity mechanisms.[65] In these models, LH receptor stimulation improves neuronal resilience and synaptic remodeling, suggesting a role in maintaining brain plasticity under conditions of hormonal fluctuation.[66] These findings indicate that physiological LH levels contribute to neuroprotection by bolstering survival pathways in key cognitive regions like the hippocampus.LH signaling has been implicated in Alzheimer's disease (AD) pathology, where elevated levels correlate with increased amyloid-beta (Aβ) accumulation. Animal studies from the early 2000s demonstrated that higher serum LH concentrations are associated with greater plasma Aβ peptide levels, independent of testosterone.[67] In transgenic mouse models of AD, genetic ablation of the LHCGR reduces Aβ deposition and tauphosphorylation, underscoring LH's potential to exacerbate amyloidpathology when dysregulated.[68]Research through the 2020s has further linked chronic LH elevation to impaired cognition and AD progression, with direct CNS exposure to LH increasing cerebral Aβ levels in guinea pigs.[69] These observations suggest that LH modulates Aβ processing via hippocampal and cortical receptors, contributing to neurodegenerative risk.In the hypothalamus, LH expression may play a regulatory role in gonadotropin dynamics, influencing local neuronal circuits. Higher LH levels in this region correlate with feedback mechanisms that could modulate the expression and release of gonadotropins through pituitary-brain interactions, though the precise pathways remain under investigation.[65]Recent research up to 2025 has explored LH's involvement in mood disorders and cognitive aging, particularly in females. In aging mouse models, absent LH signaling rescues anxiety-like behaviors, linking elevated LH to mood dysregulation during menopause transition.[70] Studies indicate that reducing LH levels alleviates depressive symptoms and cognitive deficits, with LH implicated in hypothalamic-pituitary disruptions that affect emotional processing and age-related memory decline.[71] These findings highlight LH's emerging role in non-reproductive brain health, with potential therapeutic implications for targeting CNS LH signaling in late-life mood and cognition disorders.[72]
Clinical Measurement
Normal Serum Levels
Luteinizing hormone (LH) concentrations in serum are typically quantified in international units per liter (IU/L). Due to the pulsatile secretion pattern of LH, characterized by episodic releases every 60–120 minutes driven by hypothalamic gonadotropin-releasing hormone, single measurements may not fully capture dynamic levels, often requiring serial sampling over several hours for comprehensive evaluation.[2][73][74]In females of reproductive age, LH levels fluctuate significantly with the menstrual cycle phases. During the follicular phase, baseline concentrations range from 2–10 IU/L; they increase to 20–100 IU/L during the mid-cycle ovulatory surge, which triggers final follicular maturation and ovulation; and then decline to 1–10 IU/L in the luteal phase. Postmenopausal women experience chronically elevated LH due to reduced ovarian feedback, with levels typically exceeding 20 IU/L and often ranging from 19–100 IU/L.[2][73]
In adult males, LH levels remain relatively constant at 1–10 IU/L, supporting steady testosterone production.[2][73][75]Across the lifespan, LH exhibits distinct age-related patterns: prepubertal children maintain low levels below 1 IU/L (often <0.4 IU/L), reflecting hypothalamic-pituitary-gonadal axis quiescence, followed by a progressive rise starting in early puberty (around ages 8–13 in girls and 9–14 in boys) that reaches adult ranges by mid-to-late adolescence.[76][77]Several factors can influence measured LH levels, including diurnal variation—minimal in adults but featuring augmented nocturnal pulses during puberty—and differences in assay methodologies. Immunoassays, such as immunoradiometric assays (IRMA) or enzyme-linked immunosorbent assays (ELISA), are the standard for clinical measurement due to their sensitivity and specificity for intact LH, while bioassays (e.g., rat Leydig cell or granulosa cell-based) assess biological activity but are less commonly used owing to complexity.[78][79]
Application in Ovulation Prediction
Luteinizing hormone (LH) serves as a key biomarker for predicting ovulation due to its characteristic surge, which typically rises 24-36 hours prior to the release of the egg from the ovary. This surge is detectable through both urine and serum tests, allowing for timely identification of the fertile window in reproductive health monitoring. The LH surge triggers final follicular maturation and ovulation, making it a reliable indicator when measured appropriately.[80]Home ovulation predictor kits (OPKs) are widely used for non-invasive detection of the urinary LH surge, with most kits calibrated to a sensitivity threshold of approximately 20-30 IU/L to identify the onset of the surge. These kits enable individuals to perform daily urine tests starting around cycle day 10, providing a positive result when LH levels exceed the threshold, signaling imminent ovulation. In clinical studies, OPKs have demonstrated high accuracy, predicting the LH surge within one day in 82-88% of cases and within two days in 89-96% of cases when compared to serum measurements.[81][82][83]In clinical settings, fertility monitoring protocols often involve serial blood draws to quantify serum LH levels alongside transvaginal ultrasound to assess follicular development and confirm ovulation timing. This combined approach is standard in assisted reproductive technologies, such as intrauterine insemination or in vitro fertilization, where precise LH tracking helps optimize intervention timing. Overall, LH-based methods achieve 90-95% accuracy in delineating the fertile window, though they may be less reliable in anovulatory cycles where no surge occurs, potentially leading to false negatives.[84][85][86]As of 2025, integration of LH tracking with mobile apps and wearable devices has enhanced accessibility and precision in ovulation prediction. Devices like the Inito fertility monitor pair urine test strips with smartphone apps to analyze LH alongside other hormones such as estrogen metabolites, offering quantitative data and personalized fertile window forecasts. Wearables and apps, including those from Flo and Premom, incorporate LH test results with basal body temperature and cycle data for real-time predictions, improving user compliance and conception rates in natural family planning.[87][88][89]
Pathophysiology
LH Excess Conditions
Luteinizing hormone (LH) excess refers to conditions where serum LH levels are abnormally elevated, often disrupting normal gonadal function and leading to hyperandrogenism or premature sexual development. This hypersecretion can arise from disruptions in the hypothalamic-pituitary-gonadal axis, including altered gonadotropin-releasing hormone (GnRH) pulsatility that preferentially stimulates LH release over follicle-stimulating hormone (FSH).[90]One primary cause of LH excess is polycystic ovary syndrome (PCOS), a common endocrine disorder affecting reproductive-aged women, characterized by an elevated LH/FSH ratio often exceeding 2:1. This imbalance stems from increased GnRH pulse frequency, which drives excessive LH secretion and subsequent ovarian androgen overproduction by theca cells.[52][91][92]Central precocious puberty (CPP), whether idiopathic or due to central nervous system lesions, represents another key condition involving LH excess, particularly in children. In CPP, premature activation of the hypothalamic GnRH neurons leads to pulsatile LH release, stimulating early gonadal maturation and sex steroid production.[93][94]During the menopause transition, ovarian follicle depletion results in diminished negative feedback from estrogen and inhibin, causing a natural rise in LH levels as the pituitary compensates for declining ovarian function. This elevation persists post-menopause, contributing to the hormonal profile of reproductive senescence.[95][96]Elevated LH levels also characterize hypergonadotropic hypogonadism, resulting from primary gonadal failure and reduced negative feedback from sex steroids. In males, Klinefelter syndrome (47,XXY karyotype) leads to compensatory LH elevation with low testosterone, causing symptoms such as small testes, infertility, gynecomastia, and reduced secondary sexual characteristics. In females, premature ovarian insufficiency (POI) or Turner syndrome (45,X) results in high LH and FSH levels, amenorrhea, and estrogen deficiency, often with streak gonads and infertility.[97][98]Symptoms of LH excess vary by sex and underlying cause but commonly include manifestations of androgen excess. In females, particularly those with PCOS, elevated LH contributes to irregular menstrual cycles, oligomenorrhea, and hirsutism due to increased ovarian androgen synthesis. In males, LH excess in conditions like CPP leads to precocious puberty features such as rapid growth, pubic hair development, penile and testicular enlargement, acne, and voice deepening.[91][93]
LH Deficiency Conditions
Luteinizing hormone (LH) deficiency, a key feature of hypogonadotropic hypogonadism (HH), arises from impaired hypothalamic or pituitary function, leading to insufficient gonadotropin secretion and subsequent gonadal dysfunction.[98] This condition disrupts the hypothalamic-pituitary-gonadal axis, resulting in low LH levels that fail to stimulate adequate sex hormone production in the gonads.[99]HH is classified as secondary hypogonadism, distinguishing it from primary gonadal failure where LH levels would be elevated.[98]Congenital forms of HH include Kallmann syndrome, characterized by GnRH deficiency combined with anosmia due to failed neuronal migration during embryonic development, and idiopathic HH without olfactory deficits.[100]Kallmann syndrome often involves genetic mutations, such as in the FGFR1 gene, which accounts for approximately 10% of cases and impairs fibroblast growth factor signaling essential for GnRH neuron development.[101] Idiopathic HH, lacking identifiable structural or olfactory abnormalities, represents a heterogeneous group where genetic factors may still play a role but remain undetected in many patients.[99]Secondary causes of LH deficiency encompass acquired disruptions to the hypothalamus or pituitary, including pituitary tumors that compress gonadotroph cells, traumatic brain injury damaging the hypothalamic-pituitary axis, and anorexia nervosa, which suppresses GnRH pulsatility through severe nutritional deprivation and stress.[4] Pituitary adenomas, for instance, can lead to hypopituitarism by mass effect or hormonal interference, while traumatic brain injury often affects the pituitary stalk, interrupting gonadotropin-releasing hormone (GnRH) delivery.[102] In anorexia nervosa, chronic energy deficit reversibly inhibits LH secretion, mimicking HH until weight restoration.[103]Clinical manifestations of LH deficiency include delayed or absent puberty, infertility due to impaired gametogenesis, and reduced libido from hypoandrogenism in males or hypoestrogenism in females.[98] Additional symptoms encompass erectile dysfunction, decreased muscle mass, fatigue, and mood disturbances, with long-term risks such as osteoporosis from prolonged low sex steroid levels leading to reduced bone mineral density.[99] In females, amenorrhea and infertility predominate, while males may experience gynecomastia or micropenis if onset occurs prepubertally.[100]Diagnosis of LH deficiency requires demonstration of low serum LH levels alongside subnormal gonadal hormones, such as testosterone below 300 ng/dL in males or estradiol below 20 pg/mL in females, with follicle-stimulating hormone (FSH) similarly reduced to confirm central origin.[104] Levels of LH typically below 1-2 IU/L in the context of low sex steroids support the diagnosis, often verified by repeated morning measurements to account for pulsatile secretion.[99]Imaging via MRI may identify structural causes like tumors, and genetic testing is pursued for suspected congenital forms.[100]The prevalence of congenital HH, including Kallmann syndrome, is estimated at 1 in 10,000 to 50,000 individuals, with a higher incidence in males (male-to-female ratio of 4-5:1) due to X-linked forms.[100] Acquired secondary causes vary by etiology; for example, post-traumatic brain injury HH occurs in up to 30% of severe cases, while anorexia nervosa affects up to 40% of patients with the disorder.[105] Genetic mutations, such as FGFR1 in Kallmann syndrome, are identified in about 30% of congenital HH cases overall.[106]
Medical Applications
Diagnostic Uses
Luteinizing hormone (LH) measurements play a crucial role in diagnosing reproductive and endocrine disorders by assessing the functionality of the hypothalamic-pituitary-gonadal axis. Serum LH levels, often evaluated alongside follicle-stimulating hormone (FSH), provide insights into conditions such as polycystic ovary syndrome (PCOS), amenorrhea, and hypogonadism. Abnormal LH patterns help differentiate between central (hypothalamic or pituitary) and peripheral (gonadal) etiologies, guiding further clinical evaluation.[107]Although the LH/FSH ratio was historically considered suggestive of PCOS with ratios greater than 2:1, current guidelines as of 2023 do not recommend its use for diagnosis due to low sensitivity, specificity, and assay variability; diagnosis relies on hyperandrogenism, ovulatory dysfunction, and polycystic ovarian morphology. Elevated LH levels in PCOS reflect underlying neuroendocrine disturbances contributing to ovulatory dysfunction and may support differential diagnosis, particularly in adolescents when combined with clinical, ultrasound, and anti-Müllerian hormone (AMH) findings.[108][109][110][111]Stimulation tests, such as the GnRH analog challenge, are employed to evaluate pituitary reserve and confirm the integrity of gonadotropin secretion. In this test, administration of a GnRH analog stimulates LH release, with the response at 4 hours indicating pituitary gonadotrophin reserve and the 24-hour sample reflecting gonadal feedback. A robust LH increment suggests intact pituitary function, while blunted responses point to deficiencies, aiding in the diagnosis of central hypogonadism.[112][113]LH assessment is integral to evaluating amenorrhea, where low basal LH levels indicate hypothalamic or pituitary disorders, such as functional hypothalamic amenorrhea, due to insufficient GnRH drive. In contrast, elevated LH levels signal primary ovarian failure (hypergonadotropic hypogonadism), where the ovaries fail to respond adequately, leading to compensatory gonadotropin rise. This distinction directs subsequent investigations, with low LH prompting evaluation for stress, weight loss, or tumors, and high LH suggesting ovarian insufficiency.[114][115][116]Diagnostic accuracy is enhanced by integrating LH measurements with imaging and other hormones. Magnetic resonance imaging (MRI) of the pituitary is recommended when LH is low or inconsistent, to identify structural lesions like adenomas or hypoplasia. Concurrent assessment of testosterone and estradiol levels complements LH; for instance, low testosterone with inappropriately low LH suggests secondary hypogonadism, while normal or high testosterone with elevated LH indicates primary testicular failure in males. In females, low estradiol with high LH supports ovarian failure diagnosis. In line with 2023 PCOS guidelines, LH measurements support differential diagnosis but are supplemented by AMH and other markers in adolescents and adults for precise assessment.[117][98][107][111]Recent advances in automated immunoassays have improved LH detection, particularly for low levels in children, enabling earlier diagnosis of precocious puberty or delayed puberty. These immunometric assays, such as chemiluminescent platforms, offer high sensitivity and specificity, detecting subtle LH elevations that older radioimmunoassays might miss, thus facilitating precise monitoring of pubertal onset.[118][119]
Therapeutic Uses
Recombinant human luteinizing hormone (rhLH), marketed as Luveris (lutropin alfa), is indicated for controlled ovarian hyperstimulation in women with profound LH and follicle-stimulating hormone (FSH) deficiency undergoing assisted reproductive technologies such as in vitro fertilization (IVF).[120] It is administered subcutaneously at a fixed dose of 75 IU per day in conjunction with recombinant FSH (starting 75–150 IU per day, which may be adjusted based on ovarian response and monitoring of serum estradiol levels). Clinical studies have demonstrated that rhLH supplementation enhances oocyte yield and embryo quality in LH-deficient patients, leading to higher pregnancy rates compared to FSH monotherapy.[121]Human chorionic gonadotropin (hCG), which acts as a surrogate for the endogenous LH surge due to its structural similarity and ability to bind LH receptors, is widely used for final follicular maturation and ovulation induction in IVF cycles.[122] The standard dose is a single intramuscular injection of 5,000 to 10,000 IU, administered when leading follicles reach an appropriate size (typically 17-18 mm in diameter), triggering luteinization and oocyte release approximately 36 hours later.[123] This approach has been shown to achieve comparable oocyte recovery and fertilization rates to urinary hCG, with reduced risk of immunogenicity when using recombinant forms.[124]In patients with hypogonadotropic hypogonadism, pulsatile gonadotropin-releasing hormone (GnRH) pump therapy restores physiological pulsatile LH and FSH secretion, thereby inducing ovulation in women and spermatogenesis in men to facilitate fertility restoration.[125] Administered via a portable subcutaneous pump delivering GnRH pulses every 90-120 minutes, this method mimics natural hypothalamic-pituitary-gonadal axis function and has demonstrated ovulation rates exceeding 90% in responsive patients, with live birth rates of 70-80% after multiple cycles.[126] It offers a safer profile than direct gonadotropin injections by minimizing the risk of multiple pregnancies and ovarian hyperstimulation syndrome.[127]Off-label applications of LH-like therapies, particularly hCG, extend to male hypogonadism for testosterone replacement by stimulating Leydig cell production in the testes.[128] In men with secondary hypogonadism desiring fertility preservation, hCG monotherapy (typically 1,500-5,000 IU subcutaneously 2-3 times weekly) elevates serum testosterone levels to the normal range while maintaining intratesticular testosterone necessary for spermatogenesis, avoiding the suppressive effects of exogenous testosterone.[129] This approach has been effective in achieving eugonadal testosterone concentrations and improving symptoms such as fatigue and low libido without significant adverse events.[130]Recent studies (2020–2025) suggest that LH supplementation may improve outcomes, such as oocyte yield and pregnancy rates, particularly in LH-deficient or poor ovarian responders when integrated with GnRH antagonists, though benefits vary by patient group.[131]
Cellular Mechanisms
Receptor Interaction
The luteinizing hormone/choriogonadotropin receptor (LHCGR) is a G protein-coupled receptor (GPCR) characterized by a large extracellular domain responsible for ligandbinding and a seven-transmembrane helical domain that facilitates signal transduction, primarily expressed on gonadal cells such as theca and luteal cells in the ovaries and Leydig cells in the testes, with lower expression in brain vessels and microglial cells.[132][133] Luteinizing hormone (LH) binds to the extracellular domain of LHCGR with high specificity, where the beta subunit of LH primarily determines this receptor selectivity among glycoprotein hormones, while the alpha subunit contributes to overall stability.[134]Glycosylation patterns on LH, particularly sialylation levels, influence the hormone's circulatory half-life but do not directly alter binding affinity to LHCGR.[135]Upon LH binding, LHCGR undergoes a conformational change that promotes receptor dimerization and subsequent internalization via clathrin-mediated endocytosis, involving beta-arrestin recruitment to regulate signaling duration and prevent overstimulation.[136][137] This dynamic process ensures precise control of receptor availability on the cell surface, with internalized complexes often recycled or degraded depending on the ligand concentration.[138]Polymorphisms in the LHCGR gene, such as the rs2293275 variant and the 18insLQ insertion, have been associated with altered receptor responsiveness, leading to hypo-responsiveness in ovarian stimulation protocols or hyper-responsiveness in conditions like polycystic ovary syndrome, thereby influencing LH-mediated gonadal function.[139][140]
Role in Signal Transduction
Upon binding to its receptor, luteinizing hormone (LH) activates the stimulatory G protein (Gs), which in turn stimulates adenylate cyclase to increase intracellular cyclic adenosine monophosphate (cAMP) levels.[141] This elevation in cAMP serves as a key second messenger in LH-mediated signaling within gonadal cells.[142]The rise in cAMP activates protein kinase A (PKA), which phosphorylates downstream targets including the cAMP response element-binding protein (CREB).[143] Phosphorylated CREB translocates to the nucleus and promotes transcription of genes encoding steroidogenic enzymes, such as those involved in progesterone and testosterone synthesis.[144] This PKA-dependent pathway is essential for the acute and chronic regulation of steroid hormone production in response to LH stimulation.A critical aspect of LH-induced phosphorylation involves the steroidogenic acute regulatory (StAR) protein, which facilitates cholesterol transport from the outer to the inner mitochondrial membrane—a rate-limiting step in steroidogenesis.[145] LH stimulation leads to PKA-mediated phosphorylation of StAR, enhancing its activity and enabling rapid steroid hormonebiosynthesis in Leydig and theca cells.[146] This post-translational modification is indispensable for the efficient mobilization of cholesterol substrate to cytochrome P450 side-chain cleavage enzyme.[147]In addition to the Gs/cAMP pathway, LHCGR can couple to Gq/11 proteins, activating phospholipase C (PLC) to hydrolyze phosphatidylinositol 4,5-bisphosphate (PIP2) into inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DAG). IP3 triggers the release of intracellular calcium (Ca2+), which is crucial for non-steroidogenic responses such as oocyte maturation and resumption of meiosis during ovulation.[148][149]LH signaling also exhibits cross-talk with the mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK) pathways, contributing to cell proliferation and differentiation in gonadal tissues.[150] Activation of these pathways by LH occurs through transactivation of epidermal growth factor receptors or direct coupling, promoting ERK1/2 phosphorylation and subsequent regulation of genes involved in ovarian follicle maturation and Leydig cell function.[151] This integration amplifies LH's effects beyond cAMP signaling, supporting gonadal development and hormone responsiveness.[152]To prevent overstimulation, LH receptor signaling undergoes desensitization via phosphorylation by G protein-coupled receptor kinases (GRKs), which recruits β-arrestins to uncouple the receptor from Gs and promote internalization.[153] This GRK-mediated phosphorylation occurs on serine and threonine residues in the receptor's carboxyl terminus, leading to β-arrestin binding that terminates cAMP production and facilitates receptor downregulation.[154] Such mechanisms ensure temporal control of LH responses in steroidogenic cells.[155]